content
stringlengths
1
15.9M
\subsection{General Definitions for Quantum Graphs} We shall start with a few general definitions. Graphs consist of $V$ {\it vertices} connected by $B$ {\it bonds} (or {\it edges}). The {\it valency} $v_{i}$ of a vertex $i$ is the number of bonds meeting at that vertex. Associated to every graph is its {\it connectivity (adjacency) matrix} $C_{i,j}$. It is a square matrix of size $V$ whose matrix elements $C_{i,j}$ are given in the following way \begin{eqnarray} C_{i,j}=C_{j,i}=\left\{ \begin{array}{l} 1\qquad\text{if}\ i,j\ \text{ are connected} \\ 0\qquad\text{otherwise} \end{array}\right\} \qquad(i,j=1,\dots,V)\,. \label{cmat} \end{eqnarray} The valency of a vertex is given in terms of the connectivity matrix, by $v_i= \sum_{j=1}^V C_{i,j}$ and the total number of bonds is $B= {1\over 2}\sum_{i,j=1}^VC_{i,j}$. When the vertices $i$ and $j$ are connected, we shall assume that the connection is achieved by a single bond, such that multiple bonds are excluded. We denote the connecting bond by $b=[i,j]$. Note that the notation $[i,j]$ will be used whenever we do not need to specify the {\it direction} on the bond. Hence $[i,j]=[j,i]$. {\it Directed bonds} will be denoted by $(i,j)$, and we shall always use the convention that the bond is directed from the first index to the second one. To each bond $[i,j]$ we assign a length $L_{[i,j]} = L_{(i,j)} = L_{(j,i)}$. In most applications we would avoid non-generic degeneracies by assuming that the $L_{[i,j]}$ are {\it rationally independent}. The mean length is defined by $\left \langle L \right \rangle \equiv {1\over B}\sum_{b=1}^B L_b$. For the quantum description we assign to each bond $b=[i,j]$ a coordinate $x_b$ which measures distances along the bond. We may use $x_{(i,j)}$ which is defined to take the value $0$ at the vertex $i$ and the value $L_{(i,j)} \equiv L_{(j,i)}$ at the vertex $j$. We can also use $x_{(j,i)}$ which vanishes at $j$ and takes the value $L_{(i,j)}$ at $i$. The wave function $\Psi$ is a $B-$component vector and will be written as $(\Psi_{b_1}(x_{b_1})$, $\Psi_{b_2}(x_{b_2}),\dots$, $\Psi_{b_B}(x_{b_B}))^T$ where the set $\{b_i\}_{i=1}^B$ consists of all the $B$ distinct bonds on the graph. We will call $\Psi_b(x_b)$ the component of $\Psi$ on the bond $b$. The bond coordinates $x_b$ were defined above. When there is no danger of confusion, we shall use the shorthand notation $\Psi_b(x)$ for $\Psi_b(x_b)$ and it is understood that $x$ is the coordinate on the bond $b$ to which the component $\Psi_b$ refers. The Schr\"{o}dinger equation is defined on the graph in the following way \cite{A83,A94} (see also \cite{KS99} for an extensive list of references on the subject): On each bond $b$, the component $\Psi_b$ of the total wave function $\Psi$ is a solution of the one-dimensional equation \begin{eqnarray} \left(-\i\;{\text{d}/ \text{d}x_{(i,j)}} -A_{(i,j)}\right)^2\Psi_b(x_{(i,j)})=k^2\Psi_b(x_{(i,j)}) \qquad(b=[i,j])\,. \label{schrodinger} \end{eqnarray} We included a ``magnetic vector potential" $A_{(i,j)}$, with $A_{(i,j)}= -A_{(j,i)}$ which breaks time-reversal symmetry. On each of the bonds, the general solution of (\ref {schrodinger}) is a superposition of two counter-propagating waves \begin {eqnarray} \psi_{(i,j)}(x_{(i,j)}) = \exp \left (\i\[ kx_{(i,j)} + A_{(i,j)}x_{(i,j)}\]\right ) \nonumber \\ \psi_{(j,i)}(x_{(j,i)}) = \exp \left (\i\[ kx_{(j,i)} + A_{(j,i)}x_{(j,i)}\]\right )\,. \label{counterprop} \end{eqnarray} Note that the above functions are normalised to have an amplitude $1$ at the points from which they ``emerge", namely, $\psi_{(i,j)}=1$ at the vertex $i$ and $\psi_{(j,i)}=1$ at the vertex $j$. The Hilbert space of the solutions of (\ref {schrodinger}) is spanned by the set of functions defined above, such that for all $b=[i,j]$ \begin{equation} \Psi_b = a_{(i,j)} \psi_{(i,j)}(x_{(i,j)}) + a_{(j,i)} \psi_{(j,i)}(x_{(j,i)})\,. \label {aijdef} \end{equation} Thus, the yet undetermined coefficients $a_{(i,j)}$ form a $2B$-dimensional vector of complex numbers, which uniquely determines an element in the Hilbert space of solutions. This space corresponds to ``free wave" solutions since we did not yet impose any conditions which the solutions of (\ref {schrodinger}) have to satisfy at the vertices. \subsection{The Quantum Scattering Map} The {\em quantum scattering map} is a unitary transformation acting in the space of free waves, and it is defined as follows. In a first step, we prescribe at each vertex $i=1,\dots,V$ a {\it vertex scattering matrix} which is a unitary matrix of dimension $v_i$. The vertex scattering matrices may be $k$ dependent and they are denoted by $\sigma_{l,m}^{(i)} (k)$, where the indices $l,m$ take the values of the vertices which are connected to $i$, that is, $C_{i,l}=C_{i,m}=1$. The vertex scattering matrix is a property which is attributed to the vertex under consideration. It can either be derived from appropriate boundary conditions as in \cite {KS97,KS99}, or, it can be constructed to model other physical situations. The important property of $\sigma_{l,m}^{(i)} (k)$ in the present context is, that any wave which is {\it incoming} to the vertex $i$ from the bonds $(l,i)$, and which has an amplitude $1$ at the vertex, is scattered and forms {\it outgoing} waves in the bonds $(i,m)$ with amplitudes $\sigma_{l,m}^{(i)} (k)$. Now, the quantum scattering map is represented by its effect on the $2B$-dimensional vector of coefficients ${\bf a} = \left \{a_{(i,j)}\right \} $, namely, $\bf a$ is mapped to $\bf a'$ with components \begin{equation} a'_{ b'} = \sum _{b=1}^{2B}a_{b} S_{B_{{b,b'}}}\,, \label{QSM} \end{equation} where $b$ and $b'$ run over all directed bonds, and if we denote $b=(i,j)$ and $b'= (l,m)$ \begin{equation} {S_B}_{(i,j),(l,m)}(k) = \delta _{j,l} {\rm e}^{\i L_{(i,j)}(k+A_{(i,j)})}\sigma_{i,m}^{(j)} (k)\,. \label{S_Bdef} \end{equation} The effect of $S_B$ on a wave function can be intuitively understood as follows. The coefficient $a_{(i,j)}$ is the (complex) amplitude of the wave which emerges from the vertex $i$ and propagates to the vertex $j$. Once it reaches the vertex $j$, it has accumulated a phase ${\rm e}^{i L_{(i,j)}(k+A_{(i,j)})}$ and it scatters into the bonds which emanate from $j$ with an amplitude given by the appropriate vertex scattering matrix. The new amplitude $a'_{(l=j,m)}$ consists of the superposition of all the amplitudes contributed by waves which impinge on the vertex $l=j$ and then scatter. The name ``quantum scattering" map is justified by this intuitive picture. The resulting matrix $S_B$ is a $2B \times 2B$ unitary matrix. The unitarity follows simply from the unitarity of the vertex scattering matrices, and from the fact that $S_B$ has non-vanishing entries between connected directed bonds: the incoming bond aims at the vertex from which the outgoing bond emerges. The unitarity of $S_B$ implies that its spectrum is restricted to the unit circle. In this paper we shall mainly be concerned with the spectral statistics of the eigenphases, and their relation to the underlying classical dynamics on the graph. The spectral statistics will be discussed in the next chapter. We shall use the remaining part of the present chapter to clarify two important issues. We shall first show how one can use the quantum scattering map to construct the space of solutions of the Schr\"odinger operator on the graph with boundary conditions. Then, we shall introduce the classical dynamics which corresponds to the scattering map. To define the space of ``bound states" on the graph, one has to restrict the space of wave functions by imposing appropriate boundary conditions on the vertices. The boundary conditions guarantee that the resulting Schr\"odinger operator is self-adjoint. In \cite {KS97,KS99}, we described and used one particular set of boundary conditions, which ensure continuity (uniqueness) and current conservation. Here we shall use a slight generalisation, which matches well with the spirit of the present article. We shall impose the boundary conditions in terms of a consistency requirement that the coefficients $a_{(i,j)}$ have to obey. Namely, we require that the wave function (\ref {aijdef}) is {\it stationary} under the action of the quantum scattering map. In other words, the vector {\bf a} must be an eigenvector of $S_B(k)$ with a unit eigenvalue. (see also \cite{Rochus}). This requirement can be fulfilled when \begin{equation} \det (I-S_B(k)) =0\,. \label {secular} \end{equation} In \cite {KS97,KS99} we have actually derived (\ref {secular}), for the particular case in which the vertex scattering matrices where computed form a particular set of vertex boundary conditions which impose continuity and current conservation on the vertices. The resulting vertex scattering matrices read \begin{equation} \sigma _{j,j^{\prime }}^{(i)}=\left( -\delta _{j,j^{\prime }}+{\frac{ (1+\e^{-\i\omega _i})}{v_i}}\right) C_{i,j}C_{i,j^{\prime }},\medskip\ \,\,\,\omega _i=2\arctan \frac{\lambda _i}{v_ik}\,. \label{smatrix} \end{equation} Here, $0\le \lambda_i \le \infty$ are arbitrary constants. The ``Dirichlet" (``Neumann") boundary conditions correspond to $ \lambda_i =\infty \ \ (0)$, respectively. The Dirichlet case implies total reflection at the vertex, $\sigma_{j,j^{\prime }} ^{(i)}= -\delta _{j,j^{\prime }}$. For the Neumann boundary condition we have $\sigma _{j,j^{\prime }}^{(i)}=-\delta _{j,j^{\prime }}+2/{v_i}$ which is independent of $k$. For any intermediate boundary condition, the scattering matrix approaches the Neumann expression as $k \rightarrow \infty$. Note that in all non-trivial cases ($v_i> 2$), back-scattering ($j =j^{\prime}$) is singled out both in sign and in magnitude: $\sigma _{j,j }^{(i)}$ has always a negative real part, and the reflection probability $|\sigma _{j,j }^{(i)}|^2 $ approaches $1$ as the valency $v_i$ increases. One can easily check that $\sigma ^{(i)}$ is a symmetric unitary matrix, ensuring flux conservation and time reversal symmetry at the vertex. For Neumann boundary conditions $\sigma ^{(i)}$ is a real orthogonal matrix. The spectral theory of the Schr\"odinger operators on graphs can be developed using (\ref {secular}) as the starting point. In particular, the corresponding trace formula \cite {R83} can naturally be derived, and related to the underlying classical dynamics \cite {KS97,KS99}. Here, we shall study the quantum scattering map on its own right, without a particular reference to its r\^ole in the construction of the spectrum. We shall consider the ensemble of unitary, $2B\times 2B$ matrices $S_B(k)$, where $k$ is allowed to vary in a certain interval to be specified later. Our main concern will be the statistical properties of the eigenvalues of $S_B$. This will be explained in the next chapter. \subsection{The Classical Scattering Map} The last point to be introduced and discussed in the present chapter is the classical dynamics on the graph and the corresponding scattering map. We consider a classical particle which moves freely as long as it is on a bond. The vertices are singular points, and it is not possible to write down the analogue of Newton's equations at the vertices. Instead, one can employ a Liouvillian approach based on the study of the evolution of phase-space densities. This phase-space description will be constructed on a Poincar\'{e} section which is defined in the following way. Crossing of the section is registered as the particle encounters a vertex, thus the ``coordinate" on the section is the vertex label. The corresponding ``momentum" is the direction in which the particle moves when it emerges from the vertex. This is completely specified by the label of the next vertex to be encountered. In other words, \begin{equation} \left\{ \begin{array}{c} {\rm position} \\ {\rm momentum} \end{array} \right\} \Longleftrightarrow \left\{ \begin{array}{c} {\rm vertex}\text{ }{\rm index} \\ {\rm next}\text{ }{\rm index} \end{array} \right\}\,. \end{equation} The set of all possible vertices and directions is equivalent to the set of $2B$ directed bonds. The evolution on this Poincar\'{e} section is well defined once we postulate the transition probabilities $P_{j\to j^{\prime}}^{(i)}$ between the directed bonds $b=\{j,i\}$ and $b^{\prime }=\{i,j^{\prime }\}$. To make the connection with the quantum description, we adopt the quantum transition probabilities, expressed as the absolute squares of the $S_B$ matrix elements \begin{equation} P_{j\to j^{\prime }}^{(i)}=\left| \sigma_{j,j^{\prime }}^{(i)}(k)\right| ^2\,. \label{cl1} \end{equation} When the vertex scattering matrices are constructed from the standard matching conditions on the vertices (\ref {smatrix}), we get the explicit expression \begin{equation} P_{j\to j^{\prime }}^{(i)} = \left| -\delta _{j,j^{\prime }}+{\frac{(1+\e^{-\i\omega_i})}{v_i}}\right|^2 \,. \label{cl11} \end{equation} For the two extreme cases corresponding to Neumann and Dirichlet boundary conditions this results in \begin{eqnarray} P_{j\to j^{\prime }}^{(i)} &=& \left\{ \begin{array}{ll} \left( -\delta _{j,j^{\prime }}+{2/{v_i}}\right) ^2 & \text{Neumann} \cr \delta _{j,j^{\prime}} & \text{Dirichlet} \end{array} \right\}\,. \end{eqnarray} The transition probability $P_{j\to j^{\prime }}^{(i)}$ for the Dirichlet case admits the following physical interpretation. The particle is confined to the bond where it started and thus the phase space is divided into non-overlapping ergodic components ($\approx$ ``tori''). For all other boundary conditions the graph is dynamically connected. The classical Frobenius-Perron evolution operator is a $2B\times 2B$ matrix whose elements $U_{b,b^{\prime }}$ are the classical transition probabilities between the bonds $b,b^{\prime }$ \begin{equation} U_{ij,nm}=\delta_{j,n} P^{(j)}_{i\to m}\,. \label{cl3} \end{equation} $U$ does not involve any metric information on the graph, and for Dirichlet or Neumann boundary conditions $U$ is independent of $k$. This operator is the classical analogue of the quantum scattering matrix $S_B$. Usually, one ``quantises" the classical operator to generate the quantum analogue. For graphs the process is reversed, and the classical evolution is derived from the more fundamental quantum dynamics. Let $\rho _b(t), \ b=1,\dots, 2B$ denote the distribution of probabilities to occupy the directed bonds at the (topological) time $t$. This distribution will evolve after the first return to the Poincar\'e section according to \begin{equation} \rho _b(t+1)=\sum_{b^{\prime }}U_{b,b^{\prime }}\rho _{b^{\prime }}(t)\,. \label{master} \end{equation} This is a Markovian master equation which governs the evolution of the classical probability distribution. The unitarity of the graph scattering matrix $S_B$ guarantees $\sum_{b=1}^{2B}U_{b,b^{\prime }}=1$ and $0\leq U_{b,b^{\prime }}\leq 1$, such that the probability that the particle is on any of the bonds is conserved during the evolution. The spectrum of $U$ is restricted to the unit circle and its interior, and $\nu_1 = 1$ is always an eigenvalue with the corresponding eigenvector $|1\rangle = \frac 1{2B} \left( 1,1,...,1\right) ^T$. In most cases, the eigenvalue $1$ is the only eigenvalue on the unit circle. Then, the evolution is ergodic since any initial density will evolve to the eigenvector $|1\rangle $ which corresponds to a uniform distribution (equilibrium). \begin{equation} \rho (t)\ \overrightarrow{\scriptstyle t{\rightarrow \infty }}\ |1\rangle\,. \label{epart} \end{equation} The mixing rate $-\ln \left| \nu _2\right|$ at which equilibrium is approached is determined by the gap between the next largest eigenvalue $\nu_2$ and $1$. This is characteristic of a classically mixing system. However, there are some non-generic cases such as, e.g., bipartite graphs when $-1$ belongs to the spectrum. In this case the asymptotic distribution is not stationary. Nevertheless an equivalent description is possible for bipartite graphs when $U$ is replaced by $U^2$ which has then two uncoupled blocks of dimension $B$. The example that we are going to discuss in the last section will be of this type. Periodic orbits on the graph will play an important r\^ole in the sequel and we define them in the following way. An {\it orbit} on the graph is an itinerary (finite or infinite) of successively connected {\it directed} bonds $\{i_1,i_2\}, \{i_2,i_3\},\dots $ For graphs without loops or multiple bonds this is uniquely defined by the sequence of vertices $i_1,i_2, \dots$ with $i_m \in [1,V]$ and $C_{i_m,i_{m+1}} =1$ for all $m$. An orbit is {\it periodic} with period $n$ if for all $k$, $ (i_{n+k},i_{n+k+1}) =(i_k,i_{k+1})$. The {\it code} of a periodic orbit of period $n$ is the sequence of $n$ vertices $i_1,\dots,i_n$ and the orbit consists of the bonds $(i_m,i_{m+1})$ (with the identification $i_{m+n} \equiv i_{m}$). In this way, any cyclic permutation of the code defines the same periodic orbit. The periodic orbits (PO's) can be classified in the following way: \begin{itemize} \item {\it Irreducible periodic orbits} - PO's which do not intersect themselves such that any vertex label in the code can appear at most once. Since the graphs are finite, the maximum period of irreducible PO's is $V$. To each irreducible PO corresponds its time reversed partner whose code is read in the reverse order. The only PO's which are both irreducible and conjugate to itself under time reversal are the PO's of period 2. \item {\it Reducible periodic orbits} - PO's whose code is constructed by inserting the code of any number of irreducible PO's at any position which is consistent with the connectivity matrix. All the PO's of period $n >V$ are reducible. \item {\it Primitive periodic orbits} - PO's whose code cannot be written down as a repetition of a shorter code. \end{itemize} We introduced above the concept of orbits on the graph as strings of vertex labels whose ordering obeys the required connectivity. This is a finite coding which is governed by a Markovian grammar provided by the connectivity matrix. In this sense, the symbolic dynamics on the graph is Bernoulli. This property adds another piece of evidence to the assertion that the dynamics on the graph is chaotic. In particular, one can obtain the topological entropy $\Gamma$ from the symbolic code. Using the relation \begin {equation} \Gamma = \lim_{n\to \infty} {1\over n} \log{\rm tr} (C^n) \end{equation} one gets $\Gamma = \log \bar v$, where $\bar v$ is the mean valency. Of prime importance in the discussion of the relation between the classical and the quantum dynamics are the traces $u_n={\rm tr}(U^n)$ which are interpreted as the mean classical probability to perform $n$-periodic motion. Using the definition (\ref {cl3}) one can write the expression for $u_n$ as a sum over contributions of $n$-periodic orbits \begin{equation} u_n=\sum_{p\in {\cal P}_n} n_p \exp (-r \gamma_p n_p ) \,, \label{classicalsum1} \end{equation} where the sum is over the set ${\cal P}_n$ of primitive PO's whose period $n_p$ is a divisor of $n$, with $r=n/n_p$. To each primitive orbit one can assign a {\it stability factor} $\exp (-\gamma_p n_p ) $ which is accumulated as a product of the transition probabilities as the trajectory traverses its successive vertices: \begin{equation} \exp (-\gamma_p n_p) \equiv\prod _{j=1}^{n_p} P^{(i_j)}_{i_{j-1}\to i_{j+1}}\,. \label{lyapunov} \end{equation} The stability exponents $\gamma_p$ correspond to the Lyapunov exponents in periodic orbit theory. When only one eigenvalue of the classical evolution operator $U$ is on the unit circle, one has, $u_n\overrightarrow {\scriptstyle {n\rightarrow \infty }}\ 1$. This leads to a classical sum-rule \begin{equation} u_n=\sum_{p\in P_n}n_p\exp (-r \gamma_p n_p ) \ \ \overrightarrow{ \scriptstyle {n\rightarrow \infty }}\ 1\,. \label{classicalsum} \end{equation} This last relation shows again that the number of periodic orbits must increase exponentially with $n$ to balance the exponentially decreasing stability factors of the individual periodic orbits. The topological entropy can be related to the mean stability exponent through this relation. Using the expression (\ref{classicalsum1}) for $u_n$ one can easily write down the complete thermodynamic formalism for the graph. Here, we shall only quote the periodic orbit expression for the Ruelle $\zeta $ function \begin{eqnarray} \zeta _R(z) &\equiv &\left( \det (I-zU)\right) ^{-1}={\rm \exp }\left[ -{\rm tr}\left( \ln (I-zU)\right) \right] \label{cl4} \\ &=&\exp \left[ \sum_n\frac{z^n}nu_n\right] =\prod_p\frac 1{\left( 1-z^{n_p}\exp (-n_p\gamma _p)\right)}\,, \nonumber \end{eqnarray} where the product extends over all primitive periodic orbits. The above discussion of the classical dynamics on the graph shows that it bears a striking similarity to the dynamics induced by area preserving hyperbolic maps. The reason underlying this similarity is that even though the graph is a genuinely one-dimensional system, it is not simply connected, and the complex connectivity is the origin and reason for the classically chaotic dynamics. \section{The Spectral Statistics of the Quantum Scattering Map} \label {statistics} We consider the matrices $S_B$ defined in (\ref {S_Bdef}). Their spectrum consist of $2B$ points confined to the unit circle (eigenphases). Unitary matrices of this type are frequently studied since they are the quantum analogues of classical, area preserving maps. Their spectral fluctuations depend on the nature of the underlying classical dynamics \cite{S89}. The quantum analogues of classically integrable maps display Poissonian statistics while in the opposite case of classically chaotic maps, the statistics of eigenphases conform quite accurately with the results of Dyson's random matrix theory (RMT) for the {\it circular} ensembles. The ensemble of unitary matrices which will be used for the statistical study will be the set of matrices $S_B(k)$ with $k$ in the range $|k-k_0| \le \Delta_k/2$. The interval size $\Delta_k$ must be sufficiently small such that the vertex matrices do not vary appreciably when $k$ scans this range of values. Then the $k$ averaging can be performed with the vertex scattering matrices replaced by their value at $k_0$. When the vertex scattering matrices are derived from Neumann or Dirichlet boundary conditions, the averaging interval is unrestricted because the dimension of $S_B$ is independent of $k$. In any case $\Delta_k$ must be much larger than the correlation length between the matrices $S_B(k)$, which was estimated in \cite {KS99} to be inversely proportional to the width of the distribution of the bond lengths. The ensemble average with respect to $k$ will be denoted by \begin{equation} \label{ensaveS} \left \langle \ \cdot \ \right \rangle _k \equiv \frac {1}{\Delta_k} \int_{k_0-\Delta_k/2}^{k_0+\Delta_k/2}\cdot \,\, dk \,. \end{equation} Another way to generate an ensemble of matrices $S_{B}$ is to randomise the length matrix $L$ or the magnetic vector potentials $A_{(i,j)}$, while the connectivity (topology of the graph) is kept constant. In most cases, the ensembles generated in this way will be equivalent. In the last section we will also consider an {\em additional} average over the vertex scattering matrices. In the following subsections we compare statistical properties of the eigenphases $\left\{ \theta _l(k)\right\} $ of $S_B$ with the predictions of RMT \cite{M90} and with the results of periodic orbit theory for the spectral fluctuations of quantised maps \cite{BS88}. The statistical measure which we shall investigate is the spectral form factor. Explicit expressions for this quantity are given by RMT \cite{HKSSZ96}, and a semiclassical discussion can be found in \cite{keatbog,UScorr,camb}. \subsection{\bf The Form Factor} \label{the_form_factor} The matrix $S_B$ for a fixed value of $k$ is a unitary matrix with eigenvalues $\e^{\i\theta_l(k)}$. The spectral density of the eigenphases reads \begin{equation} d(\theta;k )\equiv \sum_{l=1}^{2B}\delta (\theta -\theta _l(k))=\frac{2B}{2\pi }+ \frac 1{2\pi }\sum_{n=1}^\infty\e^{-\i\theta n}{\rm tr}S^n_B(k) +{\rm c.c.}\,, \label{sms1} \end{equation} where the first term on the r.h.s.\ is the smooth density $\overline{d}=\frac{2B}{2\pi }$. The oscillatory part is a Fourier series with the coefficients ${\rm tr}S^n_B(k) $. This set of coefficients will play an important r\^ole in the following. Using the definitions (\ref{S_Bdef}) one can expand ${\rm tr}S^n_B(k) $ directly as a sum over $n-$periodic orbits on the graph \begin{equation} {\rm tr} S^n_B(k) =\sum_{p\in {\cal P}_n}n_p{\cal A}_p^r{\rm e}^{i(kl_p+ \Phi_p)r}{\rm e} ^{i \mu _p r }\,, \label{posum} \end{equation} where the sum is over the set ${\cal P}_n$ of primitive PO's whose period $n_p$ is a divisor of $n$, with $r=n/n_p$. $l_p = \sum_{b \in p} L_{b}$ is the length of the periodic orbit. $\Phi_p = \sum_{b \in p}L_b A_b$ is the ``magnetic flux" through the orbit. If all the parameters $A_b$ have the same absolute size $A$ we can write $\Phi_p = A b_p$, where $b_p$ is the directed length of the orbit. $\mu _p$ is the phase accumulated from the vertex matrix elements along the orbit, and it is the analogue of the Maslov index. For the standard vertex matrices (\ref {smatrix}) $\mu_p/\pi$ gives the number of {\em backscatterings} along $p$. The amplitudes ${\cal A}_p$ are given by \begin{equation} {\cal A}_p=\prod _{j=1}^{n_p} \left |\sigma^{(i_j)}_{i_{j-1},i{j+1}}\right| \equiv {\rm e}^{-{\frac{ \gamma _p}2}n_p}\,, \label{amplitude} \end{equation} where $i_{j}$ runs over the vertex indices of the periodic orbit, and $j$ is understood ${\rm mod}\,n_{p}$. The Lyapunov exponent $\gamma_p$ was defined in (\ref {lyapunov}). It should be mentioned that (\ref {posum}) is the building block of the periodic orbit expression for the spectral density of the graph, which can be obtained starting from the secular equation (\ref {secular}). In the quantisation of classical area preserving maps similar expressions appear as the leading semiclassical approximations. In the present context (\ref {posum}) is an identity. The two-point correlations are expressed in terms of the excess probability density $R_2(r)$ of finding two phases at a distance $r$, where $r$ is measured in units of the mean spacing ${ 2\pi \over 2B}$ \begin{equation} R_2(r;k_0)={2\over 2\pi}\sum_{n=1}^\infty \cos \left( \frac{2\pi rn}{2B}\right) \frac 1{2B}\left\langle\left| {\rm tr}S_B^n\right|^2\right\rangle_k\,\, . \label{sms3} \end{equation} The form factor \begin{equation}\label{ff} K(n/2B)={\frac 1{2B}}<|{\rm tr}S_B^n|^2>_k \end{equation} is the Fourier transform of $R_2(r,k_0)$. For a Poisson spectrum, $K(n/2B)=1$ for all $n$. RMT predicts that $K(n/2B)$, depends on the scaled time ${n/2B}$ only \cite{S89}, and explicit expressions for the orthogonal and the unitary circular ensembles are known \cite{HKSSZ96}. As was indicated above, if the vertex scattering matrices are chosen by imposing Dirichlet boundary conditions on the vertices, the classical dynamics is ``integrable". One expects therefore the spectral statistics to be Poissonian, \begin{equation} K(n/2B)= 1\qquad {\rm for\ all}\ n\ge 1 \,. \end{equation} For Dirichlet boundary conditions the vertex scattering matrices (\ref{smatrix}) couple only time reversed bonds. $S_B$ is reduced to a block diagonal form where each bond and its time reversed partner are coupled by a $2\times 2$ matrix of the form \begin{eqnarray} S^{(b)}(k,A) =\left ( {\begin{array}{ll} 0 & \e^{\i(k+A)L_b} \\ \e^{\i(k-A)L_b } & 0 \end{array}}\right ) \,. \end{eqnarray} The spectrum of each block is the pair $\pm \e^{\i kL_b}$, with the corresponding symmetric and antisymmetric eigenvectors $ {1\over \sqrt {2}}(1, \pm1)$. As a result, we get \begin{equation} K(n/2B)=1+(-1)^n \ \ \ {\rm for \ \ all} \ \ \ n\geq 1 \,. \label{poissonform} \end{equation} This deviation from the expected Poissonian result is due to the fact that the extra symmetry reduces the matrix $S_B$ further into the symmetric and antisymmetric subspaces. The spectrum in each of them is Poissonian, but when combined together, the fact that the eigenvalues in the two spectra differ only by a sign leads to the anomaly (\ref {poissonform}). \begin{figure}[htb] \centerline{\psfig{figure=V20.eps,width=15 cm,angle=270}} \vspace*{3mm} \caption{\label{v20} Form factor for a fully connected graph with $V=20$ (a) with and (b) without time-reversal symmetry. The smooth curves show the predictions of the corresponding random matrix ensembles COE and CUE, respectively. } \end{figure} Having successfully disposed of the integrable case, we address now the more general situation. In Fig.~\ref{v20} we show typical examples of form factors, computed numerically for a fully connected graph with $V= 20$. The data for Neumann boundary conditions and $A=0$ (Fig.~1(a)) or $A\ne 0$ (Fig.~1(b)) are reproduced quite well by the predictions of RMT, which are shown by the smooth lines. For this purpose, one has to scale the topological time $n$ by the corresponding ``Heisenberg time" which is the dimension of the matrix, i.e., $2B$. The deviations from the smooth curves are not statistical, and cannot be ironed out by further averaging. Rather, they are due to the fact that the graph is a dynamical system which cannot be described by RMT in all detail. To study this point in depth we shall express the form factor in terms of the PO expression (\ref{posum}). \begin{eqnarray} K(n/2B)&=&\frac 1{2B}\left\langle \left|\sum_{p\in {\cal P}_n}n_p{\cal A}_p^r\e^{\i(kl_p+A b_p+ \pi \mu_p)r}\right|^2 \right\rangle_k \label{sms5} \\ &=&\left . \frac 1{2B} \sum_{p,p'\in {\cal P}_n} n_pn_{p\prime} {\cal A}_p^r {\cal A}_{p\prime}^{r^{\prime}} \exp \left \{\i A(r b_p-r'b_{p\prime}) +i\pi (r\mu_p-r'\mu_{p\prime})\right\} \right |_{rl_p = r^{\prime}l_{p^{\prime}}}\,. \nonumber \end{eqnarray} The $k$ averaging is carried out on such a large interval that the double sum above is restricted to pairs of periodic orbits which have exactly the same length. The fact that we choose the lengths of the bonds to be rationally independent will enter the considerations which follow in a crucial way. The largest deviations between the numerical data and the predictions of RMT occur for $n=1,2$. For $n=1$ one gets $0$ instead of the COE (CUE) values $1/B$ ($1/2B$), simply because the graph has no periodic orbits of period $1$. This could be modified by allowing loops, which were excluded here from the outset. The $2$-periodic orbits are self-retracing (i.e.\ invariant under time reversal), and each has a distinct length. Their contribution is enhanced because back scattering is favoured when the valency is large. Self-retracing implies also that their contribution is insensitive to the value of $A$. The form factor for $n=2$ calculated for a fully connected graph with $v=V-1$ is \begin{equation} \label{tsampi} K(n/2B)= 2\(\[1-\frac 2v\]\)^4\,, \end{equation} independent of the value of $A$. This is different from the value expected from RMT. The repetitions of the 2-periodic orbits are also the reason for the odd-even staggering which is seen for low values of $\tau\equiv n/2B$. They contribute a term which is $\approx 2\exp(-2V\tau)$ and thus decays faster with the scaled time $\tau$ when the graph increases. The deviations between the predictions of RMT and periodic orbit theory for low values of $\tau$ are typical and express the fact that for deterministic systems in general, the short time dynamics is not fully chaotic. The short time domain becomes less prominent as $B$ becomes larger because the time $n$ has to be scaled by $2B$. This limit is the analogue of the limit $\hbar \rightarrow 0$ in a general system. Consider now the domain $2 <n \ll 2B$. The PO's are mostly of the irreducible type, and the length restriction limits the sum to pairs of orbits which are conjugate under time reversal. Neglecting the contributions from repetitions and from self-retracing orbits we get \begin{equation} \label{transuzy1} K(n/2B)\approx \frac 1{2B} \sum_{p\in {\cal P}_n} n^2 {\cal A}_p^2 \ \ 4\cos ^2A b_p = {2n\over 2B} u_n \left \langle \cos ^2 A b_p\right \rangle _n \,. \end{equation} The classical return probability $u_n$ approaches $1$ as $n$ increases (see (\ref{classicalsum})). Neglecting the short time deviations, we can replace $u_n$ by $1$, and we see that the remaining expression is the classical expectation of $\cos ^2 A b_p$ over PO's of length $n$. For $A=0$ this factor is identically $1$ and one obtains the leading term of the COE expression for $n\ll 2B$. If $A$ is sufficiently large $ \left \langle \cos ^2 A b_p \right \rangle_n \approx 1/2 $, one obtains the short-time limit of the CUE result. The transition between the two extreme situations is well described by \begin{equation} \label{transuzy2} \left \langle \cos ^2 A b_p\right \rangle _n \approx {1\over 2} \left ( \e^{- A^2\left \langle L_b^2\right \rangle {n\over 2}} +1 \right ) \,. \end{equation} This formula is derived by assuming that the total directed length $b_p$ of a periodic orbit is a sum of elementary lengths with random signs. The basic approximation so far was to neglect the interference between contributions of periodic orbits with different codes (up to time reversal). This can be justified as long as periodic orbits with different codes have different lengths. This is the case for low values of $n$. As $n$ approaches $B$ the degeneracy of the length spectrum increases, and for $n>2B$ all the orbits are degenerate. In other words, the restriction $rl_p = r^{\prime}l_{p^{\prime}}$ in (\ref {sms5}) does not pick up a unique orbit and its time reversed partner, but rather a group of {\em isometric} but distinct orbits. Therefore, the interference of the contributions from these orbits must be calculated. The relative sign of the terms is determined by the ``Maslov" index. The computation of the interfering contributions from different periodic orbits with neighbouring actions is an endemic problem in the semiclassical theory of spectral statistics. These contributions are referred to as the {\it non-diagonal} terms, and they are treated by invoking the concept of periodic orbit correlations \cite {ADDKKSS93,CPS98}. The dynamical origin of these correlations is not known. In the case of graphs, they appear as correlations of the ``Maslov" signs within a class of isometric $n$-periodic orbits. To compute $K(n/2B)$ from (\ref {sms5}) one has to sum the contributions of all the $n$-periodic orbits after grouping together those which have exactly the same lengths. We shall discuss the case $A=0$, so a further restriction on the orbits to have the same directed length is not required here. Since the lengths of the individual bonds are assumed to be rationally independent, a group of isometric $n$-periodic orbits is identified by the non-negative integers $q_i, i=1,\dots,B$ such that \begin{equation} l_{\bf q} \equiv \sum_{i=1} ^B q_i l_i \qquad{\rm with}\qquad\sum_{i=1}^Bq_i=n\,, \label{qdef} \end{equation} i.e., each bond $i$ is traversed $q_i$ times. The orbits in the group differ only in the {\it order} by which the bonds are traversed. We shall denote the number of isometric periodic orbits by $D_{n}(\bf q)$. Note that not all the integer vectors ${\bf q}$ which satisfy (\ref {qdef}) correspond to periodic orbits. Rather, the connectivity required by the concept of an orbit imposes restrictions, which render the problem of computing $D_n({\bf q})$ a very hard combinatorial problem \cite {Urigavish}. Writing (\ref {sms5}) explicitly for the case of a fully connected graph with Neumann vertex scattering matrices, we get \begin{equation} K(n/2B)={1\over 2B}\left({2\over v}\right)^{2n} \sum_{\bf q} \left|\sum_{\alpha=1}^{D_n(\bf q)} {n \over r_{\alpha}} (- \xi )^{\mu_{\alpha}}\right| ^2 \ , \ \ {\rm with} \ \ \ \xi \equiv \left({v-2\over 2}\right)\,, \label {tracecomp} \end{equation} and the $\alpha$ summation extends over the $n$-periodic orbits in the class ${\bf q}$. $\mu_{\alpha}$ is the number of back scattering along the orbit, and $r_{\alpha}$ is different from unity if the orbit is a repetition of a shorter primitive orbit of period $n/r_{\alpha}$. Equation (\ref {tracecomp}) is the starting point of the new approach to spectral statistics, which we would like to develop in the present paper. The actual computation of (\ref {tracecomp}) can be considered as a {\em combinatorial} problem, since it involves counting of loops on a graph, and adding them with appropriate (signed) weights. For Neumann boundary conditions, the weights are entirely determined by the connectivity of the graph. Our numerical data convincingly show that in the limit of large $B$ the form factors for sufficiently connected graphs reproduce the results of RMT. The question is, if this relation can be derived using asymptotic combinatorial theory. The answer is not yet known, but we would like to show in the next section that for a very simple graph one can use combinatorics to evaluate the periodic orbit sums, and recover in this way the exact values of the form factor. \section{\bf The $2$-star Model} In this section we will investigate the classical and quantum dynamics in a very simple graph using two different methods. We shall use periodic orbit theory to reduce the computation of the trace of the classical evolution operator $u_{n}$ and the spectral form factor $K(n/2B)$ to combinatorial problems, namely sums over products of binomial coefficients. The result will be compared to a straight forward computation starting from the eigenvalues of the classical and quantum scattering maps. An $n$-star graph consists of a ``central" vertex (with vertex index $o$) out of which emerge $n$ bonds, all terminating at vertices (with indices $j=1,\dots, n$) with valencies $v_j=1$. The bond lengths are $L_{oj}\equiv L_j$. This simple model (sometimes called a {\it hydra}) was studied at some length in \cite {KS99}. The star with $n=2$ is not completely trivial if the central vertex scattering matrix is chosen as \begin {equation} \sigma^{(o)}(\eta ) = \left ( {\begin{array} {ll} \cos \eta & {\rm i}\sin \eta \\ {\rm i}\sin \eta & \cos \eta \end{array}} \right )\,, \label {2-starsigma} \end{equation} where the value $0\le \eta \le \pi /2 $ is still to be fixed. The scattering matrices at the two other vertices are taken to be $1$ and correspond to Neumann boundary conditions. The dimension of $U$ and $S_B$ is $4$, but it can be immediately reduced to $2$: due to the trivial scattering at the reflecting tips, $a_{jo}=a_{oj}\equiv a_j$ for $j=1,2$. In this representation the space is labelled by the indices of the two loops (of lengths $2L_1$ and $2L_2$ respectively) which start and end at the central vertex. After this simplification the matrix $S_B$ reads \begin {equation} S_B(k;\eta) = \left ( {\begin{array} {ll} {\rm e}^{2ikL_1} & 0 \\ 0 & {\rm e}^{2ikL_2} \end{array}} \right ) \left ( {\begin{array} {ll} \cos \eta & {\rm i}\sin \eta \\ {\rm i}\sin \eta & \cos \eta \end{array}} \right )\,. \label {2-starSB} \end{equation} We shall compute the form-factor for two ensembles. The first is defined by a fixed value of $\eta = \pi/4$, and the average is over an infinitely large $k$ range. The second ensemble includes an additional averaging over the parameter $\eta$. We will show that the measure for the integration over $\eta$ can be chosen such that the model yields the CUE form factor. This is surprising at first sight, since the model defined above is clearly time-reversal invariant. However, if we replace $kL_1$ and $kL_2$ in (\ref{2-starSB}) by $L(k\pm A)$, (\ref{2-starSB}) will allow for an interpretation as the quantum scattering map of a graph with a single loop of length $L$ and a vector potential $A$, i.e., of a system with broken time-reversal invariance (see Fig.~\ref{tst}). In particular, the form factors of the two systems will coincide exactly, when an ensemble average over $L$ is performed. Clearly, this is a very special feature of the model considered, and we will not discuss it here in more detail. \subsection {\bf Periodic Orbit Representation of $u_n$} The classical evolution operator corresponding to (\ref{2-starSB}) is \begin {equation} U(\eta) = \left ( {\begin{array} {ll} \cos ^2\eta & \sin ^2 \eta \\ \sin ^2\eta & \cos ^2 \eta \end{array}} \right )\,. \label {2-starclass} \end{equation} The spectrum of $U$ consists of $\{1,\cos 2\eta \}$, such that \begin{equation}\label{un} u_n (\eta )=1+\cos ^n 2\eta\,. \end{equation} We will now show how this result can be obtained from a sum over the periodic orbits of the system, grouped into classes of isometric orbits. This grouping is not really necessary for a classical calculation, but we would like to stress the analogy to the quantum case considered below. The periodic orbits are uniquely encoded by the loop indices, such that each $n$-tuple of two symbols $1$ and $2$ corresponds (up to a cyclic permutation) to a single periodic orbit. When $n$ is prime, the number of different periodic orbits is $N_2(n)=2+(2^n-2)/n$, otherwise there are small corrections due to the repetitions of shorter orbits. These corrections are the reason why it is more convenient to represent a sum over periodic orbits of length $n$ as a sum over all possible code words, though some of these code words are related by a cyclic permutation and consequently denote the same orbit. If we do so and moreover replace the stability factor of each orbit by (\ref{lyapunov}), the periodic orbit expansion of the classical return probability becomes \begin{eqnarray}\label{un_po} u_{n}&=& \sum_{i_{1}=1,2} \dots \sum_{i_{n}=1,2}\prod_{j=1}^{n} P_{i_{j}\rightarrow i_{j+1}}\,, \end{eqnarray} where $j$ is a cyclic variable such that $i_{n+1}\equiv i_{1}$. In fact (\ref{un_po}) can be obtained without any reference to periodic orbits if one expands the intermediate matrix products contained in $u_{n}=\Tr U^{n}$ and uses $P_{i_{j}\rightarrow i_{j+1}}=U_{i_{j},i_{j+1}}(\eta)$. We will now order the terms in the multiple sum above according to the classes of isometric orbits. In the present case a class is completely specified by the integer $q\equiv q_1$ which counts the traversals of the loop $1$, i.e., the number of symbols $1$ in the code word. Each of the $q$ symbols $1$ in the code is followed by an uninterrupted sequence of $t_{j}\ge 0$ symbols $2$ with the restriction that the total number of symbols $2$ is given by \begin{equation} \sum_{j=1}^{q}t_{j}=n-q\,. \end{equation} We conclude that each code word in a class $0<q<n$ which starts with a symbol $i_{1}=1$ corresponds to an ordered partition of the number $n-q$ into $q$ non-negative integers, while the words starting with $i_{1}=2$ can be viewed as partition of $q$ into $n-q$ summands. To make this step very clear, consider the following example: All code words of length $n=5$ in the class $q=2$ are $11222$, $12122$, $12212$, $12221$ and $22211$, $22121$, $21221$, $22112$, $21212$, $21122$. The first four words correspond to the partitions $0+3=1+2=2+1=3+0$ of $n-q=3$ into $q=2$ terms, while the remaining $5$ words correspond to $2=0+0+2=0+1+1=1+0+1=0+2+0=1+1+0=2+0+0$. In the multiple products in (\ref{un_po}), a forward scattering along the orbit is expressed by two different consecutive symbols $i_{j}\ne i_{j+1}$ in the code and leads to a factor $\sin^2\eta$, while a back scattering contributes a factor $\cos^2\eta$ . Since the sum is over periodic orbits, the number of forward scatterings is always even and we denote it with $2\nu$. It is then easy to see that $\nu$ corresponds to the number of positive terms in the partitions introduced above, since each such term corresponds to an uninterrupted sequence of symbols $2$ enclosed between two symbols $1$ or vice versa and thus contributes two forward scatterings. For the codes starting with a symbol $1$ there are ${q\choose \nu}$ ways to choose the $\nu$ positive terms in the sum of $q$ terms, and there are ${n-q-1\choose \nu-1}$ ways to decompose $n-q$ into $\nu$ {\em positive} summands. After similar reasoning for the codes starting with the symbol $2$ we find for the periodic orbit expansion of the classical return probability \begin{eqnarray}\label{un_ex} u_{n}(\eta)&=&2\cos^{2n}\eta+\sum_{q=1}^{n-1}\sum_{\nu} \[{q\choose \nu}{n-q-1\choose \nu-1}+{n-q\choose \nu}{q-1\choose \nu-1}\] \sin^{4\nu}\!\eta\,\cos^{2n-4\nu}\!\eta \nonumber \\ &=&2\cos^{2n}\eta+\sum_{q=1}^{n-1}\sum_{\nu}{n\/\nu} {q-1\choose \nu-1}{n-q-1\choose \nu-1}\sin^{4\nu}\!\eta\,\cos^{2n-4\nu}\!\eta\, \nonumber \\ &=&2\sum_{\nu}{n\choose 2\nu}\sin^{4\nu}\!\eta\,\cos^{2n-4\nu}\!\eta \nonumber \\ &=&(\cos^{2}\!\eta+\sin^{2}\!\eta)^{n}+(\cos^{2}\!\eta-\sin^{2}\!\eta)^{n}\,, \end{eqnarray} which is obviously equivalent to (\ref{un}). The summation limits for the variable $\nu$ are implicit since all terms outside vanish due to the properties of the binomial coefficients. In order to get to the third line we have used the identity \begin{equation}\label{ci1} \sum_{q=1}^{n-1}{q-1\choose \nu-1}{n-q-1\choose \nu-1}= {n-1\choose 2\nu-1} ={2\nu\over n}{n\choose 2\nu}\,. \end{equation} It can be derived by some straightforward variable substitutions from \begin{equation} \sum_{k=l}^{n-m}{k\choose l}{n-k\choose m}={n+1\choose l+m+1}\,. \end{equation} which, in turn, is found in the literature \cite{prudnikov}. \subsection {\bf Quantum Mechanics: Spacing Distribution and Form Factor}\label{qm} Starting from (\ref {2-starSB}), and writing the eigenvalues as ${\rm e}^{ik(L_1+L_2)} {\rm e}^{\pm i\lambda/2}$, we get for $\lambda$, the difference between the eigenphases, \begin{equation} \lambda = 2\,{\rm arcos}\[\cos\eta\,\cos k(L_1-L_2)\]\,. \label {2-starlambda} \end{equation} For fixed $\eta$, the $k$ averaged spacing distribution (which is essentially equivalent to $R_{2}(r)$ for the considered model) is given by \begin{eqnarray} \label{2-starspacing} P(\theta ;\eta) &=& {1\over \Delta_k} \int _{k_0-\Delta_k/2}^{k_0+\Delta_k/2} {\rm d}k \ \delta \left (\theta - 2 {\rm arcos} \left [\cos \eta \cos k(L_1-L_2) \right ] \right ) \nonumber \\ \nonumber \\ &=& \left \{ {\begin {array} {cl} 0 & \qquad\cos(\theta/2)> |\cos \eta\,| \\ \\ {\displaystyle\sin(\theta/2) \over \sqrt{\displaystyle\cos^2\eta-\cos^2(\theta/2)}} & \qquad\cos(\theta/2) < |\cos\eta\,| \end {array}} \right. \end{eqnarray} We have assumed that $\theta$ is the smaller of the intervals between the two eigenphases, i.e. $0\le \theta \le \pi$. The spacings are excluded from a domain centered about $0$ $(\pi)$, i.e., they show very strong level repulsion. The distribution is square-root singular at the limits of the allowed domain. $P(\theta ;\eta)$ can be written as \begin {equation} P(\theta;\eta ) = {1\over 2\pi} + {1\over \pi}\sum_{n=1}^\infty \cos(n\theta)\,\( {1\over 2}\left\langle\left|{\rm tr}S_B(\eta) ^n\right|^2\right\rangle_k -1\)\,, \label{2-starPassum} \end{equation} and, by a Fourier transformation, we can compute the form factor \begin{equation} K_{2}(n;\eta)={1\over2} \left\langle\left|{\rm tr}S_B(\eta) ^n\right|^2\right\rangle\,. \end{equation} In particular, for $\eta =\pi/4$ one finds \begin{eqnarray}\label{K2PI4_UZY} K_{2}(n;\pi/4)&=&1+{(-1)^{m+n}\over 2^{2m+1}}{2m\choose m} \\ &\approx& 1 + {(-1)^{m+n}\over 2\sqrt{\pi n }}\,. \label{k2pi4_uzy_app} \end{eqnarray} Where $ m =[n/2]$ and $[\cdot]$ stands for the integer part. The slow convergence of $K_{2}(n;\pi/4)$ to the asymptotic value $1$ is a consequence of the singularity of $P(\theta;\pi /4 )$. We now consider the ensemble for which the parameter $\eta$ is distributed with the measure ${\rm d}\mu(\eta) = |\cos\eta \sin\eta |{\rm d}\eta$. The {\em only} reason for the choice of this measure is that upon integrating (\ref {2-starPassum}) one gets \begin {equation} P(\theta) = 2\sin^2(\theta/2)\,, \label {2-starCUE} \end{equation} which coincides with the CUE result for $2\times 2$ matrices. A Fourier transformation results in \begin {equation} K_2(n) = \left \{ {\begin {array} {ll} {1\over 2} & {\rm for} \ \ n=1 \\ 1 & {\rm for} \ \ n\ge 2 \end {array} } \right.\,. \label {2-starK(n)CUE} \end {equation} The form factors (\ref{K2PI4_UZY}), (\ref{k2pi4_uzy_app}) and (\ref{2-starK(n)CUE}) are displayed in Fig.~\ref{tst} below. \subsection{Periodic Orbit Expansion of the Form Factor}\label{po} As pointed out at the end of section \ref{the_form_factor}, the $k$-averaged form factor can be expressed as a sum over classes of isometric periodic orbits. The analogue of (\ref{tracecomp}) for the 2-star is \begin{equation} K_2(n;\eta)={1\over 2}\sum_{q=0}^{n} \left|\sum_{\alpha=1}^{D_{n}(q)}{n\over r_{\alpha}} \i^{2\nu_{\alpha}}\sin^{2\nu_{\alpha}}\!\eta\cos^{n-2\nu_{\alpha}}\!\eta \right|^{2}\,, \end{equation} where the number of forward and backward scatterings along the orbits are $2\nu_\alpha$ and $\mu_{\alpha}=n-2\nu_{\alpha}$, respectively. Again, it is very inconvenient to work with the repetition number $r_{\alpha}$, and consequently we replace---as in the derivation of (\ref{un_ex})---the sum over orbits by a sum over all code words and use the analogy with the compositions of integer numbers to obtain \begin{eqnarray}\label{K2eta} K_2(n;\eta)&=& \cos^{2n}\!\eta+{n^2\over 2}\sum_{q=1}^{n-1} \[\sum_{\nu}{(-1)^{\nu}\/\nu}{q-1\choose \nu-1}{n-q-1\choose \nu-1} \sin^{2\nu}\!\eta\,\cos^{n-2\nu}\!\eta\]^{2}\,. \end{eqnarray} The inner sum over $\nu$ can be written in terms of Krawtchouk polynomials \cite{kp1,kp2} as \begin{eqnarray}\label{K2eta_K} K_2(n;\eta)&=& \cos^{2n}\!\eta+{1\/2}\sum_{q=1}^{n-1} {n-1\choose n-q}\cos^{2q}\!\eta\sin^{2(n-q)}\!\eta \[{n\over q}P_{n-1,n-q}^{(\cos^{2}\!\eta,\sin^2\!\eta)}(q)\]^2\,, \end{eqnarray} and the Krawtchouk polynomials are defined as in \cite{kp1,kp2} by \begin{eqnarray}\label{krawtchouk} P_{N,k}^{(u,v)}(x)=\[{N\choose k}(uv)^{k}\]^{-1/2}\sum_{\nu=0}^{k} (-1)^{k-\nu}{x\choose \nu}{N-x\choose k-\nu}u^{k-\nu}v^{\nu}\qquad \(\begin{array}{l}0\le k \le N\cr u+v=1\end{array}\)\,. \end{eqnarray} These functions form a complete system of orthogonal polynomials of integer $x$ with $0\le x\le N$. They have quite diverse applications ranging from the theory of covering codes \cite {cohen} to the statistical mechanics of polymers \cite{schulten}, and are studied extensively in the mathematical literature \cite{kp1,kp2}. The same functions appear also as a building block in our periodic orbit theory of Anderson localisation on graphs \cite {HSUS}. Unfortunately, we were not able to reduce the above expression any further by using the known sum-rules and asymptotic representations for Krawtchouk polynomials. The main obstacle stems from the fact that in our case the three numbers $N,k,x$ in the definition (\ref{krawtchouk}) are constrained by $N=k+x-1$. We will now consider the special case $\eta=\pi/4$ for which we obtained in the previous subsection the solution (\ref{K2PI4_UZY}). The result can be expressed in terms of Krawtchouk polynomials with $u=v=1/2$ which is also the most important case for the applications mentioned above. We adopt the common practice to omit the superscript $(u,v)$ in this special case and find \begin{eqnarray}\label{k2pi4} K_2(n;\pi/4)&=& {1\over 2^{n}}+{1\/2^{n+1}}\sum_{q=1}^{n-1} {n-1\choose n-q}\[{n\over q}P_{n-1,n-q}(q)\right]^2\,. \end{eqnarray} It is convenient to introduce \begin{eqnarray}\label{} \nq(s,t)&=&(-1)^{s+t}{s+t-1\choose s}^{1/2}P_{s+t-1,s}(t) \nonumber\\ &=&\sum_{\nu}(-1)^{t-\nu}{t\choose \nu}{s-1\choose \nu-1} \end{eqnarray} and to rewrite (\ref{k2pi4}) with the help of some standard transformations of binomial coefficients as \begin{eqnarray}\label{K2PI4N} K_2(n;\pi/4)&=&{1\over 2^{n}}+{1\/2^{n+1}}\sum_{q=1}^{n-1} \[{n\over q}\nq(q,n-q-1)\right]^2 \nonumber\\ &=&{1\over 2^{n}}+{1\over 2^{n+1}}\sum_{q=1}^{n-1}\[\nq(q,n-q)+(-1)^{n}\nq(n-q,q)\]^{2} \end{eqnarray} This expression is displayed in Fig.~\ref{tst} together with (\ref{K2PI4_UZY}) in order to illustrate the equivalence of the two results. An independent proof for this equivalence can be given by comparing the generating functions of $K_2(n;\pi/4)$ in the two representations \cite{gregory}. We defer this to appendix \ref{proof}. \begin{figure}[tb] \centerline{\psfig{figure=tst_gue.eps,width=15cm,angle=270}} \caption{\label{tst} Form factor for the 2-star quantum graph. The crosses and the connecting heavy full line show the two equivalent exact results (\protect\ref{K2PI4_UZY}) and (\protect\ref{k2pi4}) for $\eta=\pi/4$. The thin dashed lines represent the approximation (\protect\ref{k2pi4_uzy_app}), and the thin straight line corresponds to the diagonal approximation, when repetitions of primitive periodic orbits are neglected. The heavy dashed line exhibits the form factor of a CUE ensemble of $2\times 2$ random matrices (\protect\ref{2-starK(n)CUE}), which can be obtained from the 2-star by an appropriate averaging over $\eta$. Finally, the inset shows a sketch of the two possible realisations of the system: a time-reversal invariant 2-star with bond lengths $L_{1}, L_{2}$ or a graph with a single loop of length $L$ and a magnetic flux $A$ breaking time-reversal symmetry.} \end{figure} Please note, that in this way we have found a proof for two identities involving Krawtchouk polynomials \begin{equation}\label{ci2a} \sum_{q=1}^{2m-1}{2m-1\choose 2m-q}\[{2m\over q}P_{2m-1,2m-q}(q)\]^{2} =2^{2m+1}+(-1)^{m}{2m\choose m}-2 \end{equation} and \begin{equation}\label{ci2b} \sum_{q=1}^{2m}{2m\choose 2m+1-q}\[{2m+1\over q}P_{2m,2m+1-q}(q)\]^{2} =2^{2m+2}-2\,(-1)^{m}{2m\choose m}-2\,, \end{equation} which were obtained by separating even and odd powers of $n$ in (\ref{K2PI4_UZY}) and (\ref{k2pi4}). To the best of our knowledge, (\ref{ci2a}) and (\ref{ci2b}) were derived here for the first time. Finally we will derive the CUE result (\ref{2-starK(n)CUE}) for the ensemble of graphs defined in the previous subsection starting from the periodic orbit expansion (\ref{K2eta}). We find \begin{eqnarray} K_{2}(n)&=&\int_{0}^{\pi/2}{\rm d}\mu(\eta) K_2(n;\eta)\,. \end{eqnarray} Inserting (\ref{K2eta}), expanding into a double sum and using \begin{equation} \int_{0}^{\pi/2}{\rm d}\eta \sin^{2(\nu+\nu')+1}\!\eta\cos^{2(n-\nu-\nu')+1}\!\eta= {1\over 2(n+1)}{n\choose \nu+\nu'}^{-1} \end{equation} we get \begin{eqnarray}\label{intermediate} K_{2}(n)&=&{1\/n+1}+ \\\nonumber &&+ {n^2\over 4(n+1)}\sum_{q=1}^{n-1} \sum_{\nu,\nu'} {(-1)^{\nu+\nu'}\/\nu\nu'}{n\choose \nu+\nu'}^{-1} {q-1\choose \nu-1}{n-q-1\choose \nu-1}{q-1\choose \nu'-1}{n-q-1\choose \nu'-1}\,. \end{eqnarray} Comparing this to the equivalent result (\ref{2-starK(n)CUE}) we were again led to a previously unknown identity involving a multiple sum over binomial coefficients. It can be expressed as \begin{equation}\label{ci3} S(n,q)=\sum_{\nu,\nu'}F_{\nu,\nu'}(n,q)=1\qquad (1\le q < n) \end{equation} with \begin{eqnarray} F_{\nu,\nu'}(n,q)&=&{(n-1)n\/2}{(-1)^{\nu+\nu'} \over \nu\nu'} {n\choose \nu+\nu'}^{-1} {q-1\choose \nu-1}{q-1\choose \nu'-1}{n-q-1\choose \nu-1}{n-q-1\choose \nu'-1}\,. \end{eqnarray} In this case, an independent computer-generated proof was found \cite{akalu}, which is based on the recursion relation \begin{equation}\label{recursion} q^2F_{\nu,\nu'}(n,q)-(n-q-1)^{2}F_{\nu,\nu'}(n,q+1)+(n-1)(n-2q-1)F_{\nu,\nu'}(n+ 1,q+1)=0\,. \end{equation} This recursion relation was obtained with the help of a Mathematica routine \cite{multisum}, but it can be checked manually in a straight forward calculation. By summing (\ref{recursion}) over the indices $\nu,\nu'$, the same recursion relation is shown to be valid for $S(n,q)$ \cite{multisum,A=B} and the proof is completed by demonstrating the validity of (\ref{ci3}) for a few initial values. Having proven (\ref{ci3}) we can use it to perform the summation over $\nu,\nu'$ in (\ref{intermediate}) and find \begin{eqnarray} K_{2}(n)={1\/n+1}+\sum_{q=1}^{n-1}{n\/n^2-1}={1\/n+1}+{n\over n+1}(1-\delta_{n,1})\,, \end{eqnarray} which is now obviously equivalent to the random matrix form factor (\ref{2-starK(n)CUE}). To the best of our knowledge, this is the first instance in which a combinatorial approach to random matrix theory is employed. \section{\bf Conclusions} We have shown how within periodic orbit theory the problem of finding the form factor (the spectral two-point correlation function) for a quantum graph can be exactly reduced to a well-defined combinatorial problem. For this purpose it was necessary to go beyond the diagonal approximation and to take into account the correlations between the periodic orbits. In our model, these correlations are restricted to groups of isometric periodic orbits. This fits very well with the results of \cite{CPS98}, where for a completely different system (the Sinai billiard), the classical correlations between PO's were analysed and found to be restricted to relatively small groups of orbits. The code words of the orbits belonging to one group were conjectured to be related by a permutation and a symmetry operation, which is in complete analogy to the isometric orbits on graphs. Even for the very small and simple graph model that we considered in the last section the combinatorial problems involved were highly non-trivial. In fact we encountered previously unknown identities which we could not have obtained if it were not for the second independent method of computing the form factor. However, since the pioneering work documented in \cite{A=B} the investigation of sums of the type we encountered in this paper is a rapidly developing subject, and it can be expected that finding identities like (\ref{ci2a}), (\ref{ci2b}) and (\ref{ci3}) will shortly be a matter of computer power. The universality of the correlations between periodic orbits in all chaotic systems poses the problem to identify the common dynamical reasons for their occurrence and to find a common mathematical structure which is capable to describe them. A very interesting question in this respect is, if the correlations between PO's in a general chaotic system can be related to combinatorial problems. \section{\bf Acknowledgements} This research was supported by the Minerva Center for Physics of Nonlinear Systems, and by a grant from the Israel Science Foundation. We thank Tsampikos Kottos for preparing the data for Fig.~\ref{v20}. We were introduced to the {\it El Dorado} of combinatorial theory by Uri Gavish and we thank him as well as Brendan McKay and Herbert Wilf for their interest and support. We are indebted to Gregory Berkolaiko for his idea concerning the proof of (\ref{ci2a}) and (\ref{ci2b}), and to Akalu Tefera for his kind help in obtaining a computer-aided proof of (\ref{ci3}). HS wishes to thank the Weizmann Institute of Science for the kind hospitality during the visit where this work was initiated.
\section{Introduction} There can be no doubt: gauge field theories (for short, `gauge theories'), nowadays, provide a most powerful tool in modern physics with regard to a unification of the four known interaction forces. In this connection, the so-called gauge principle lays the foundation of these theories in terms of an elegant derivation of the interaction coupling. The principle works by satisfying a gauge postulate, the heartpiece of any gauge theory, which demands the theory's invariance under local gauge transformations of the matter fields. Unfortunately, we are far away from a proper understanding of the gauge principle's conceptual meaning -- it actually works just heuristically. But since the theoretical and, most of all, experimental success of the gauge approach is hardly understandable as pure coincidence, we are challenged with a deep physical and philosophical problem. It is well known that gauge field theories allow a natural mathematical description in the framework of fiber bundles, which may therefore be considered as an enlarged geometrical arena of physics. Thus, from the philosopher's point of view a first step into a better understanding could be made by analyzing the status of this geometry and its internal spaces. This paper will deal with these questions in terms of a confrontation of relationalism vs. substantivalism with regard to bundle spaces. First, I will consider the concepts of the gauge principle, gauge transformations, and gauge freedom. After introducing fiber bundles I propose a definition of the notion of gauge field theory. I will, finally, turn to the spacetime hole argument, and will propose a generalized bundle-space hole argument, which rules out fiber bundle substantivalism. \section{The gauge principle} \label{gaugeprin} We start from the empirical fact that there exist certain conserved quantities in nature. Actually, {\sc Noether}'s theorem\footnote{I refer to {\sc Emmy Noether}'s first theorem simply as {\sc Noether}'s theorem, whereas her second theorem, which is related to infinite symmetry transformations, i.e. transformations with arbitrary functions instead of parameters, for these purposes will be better described in terms of local gauge transformations, which in fact play the central role in gauge theories.} tells us that, given any global symmetry, there is a corresponding conserved quantity. \begin{quote} {\em {\bf {\sc Noether}'s theorem.} Let $\phi_i(x)$ be some field variable (with general index $i$ of the field components). Then the invariance of the action functional $S[\phi]=\int {\cal L} \left( \phi_i(x), \partial_\mu \phi_i(x) \right) \, d^4x$ under some $k$-dimensional {\sc Lie} group leads to the existence of $k$ conserved currents. } \end{quote} As a paradigm case I shall consider the free {\sc Dirac} field $\psi(x)$ with the Lagrangian density \begin{equation}\label{dirac_L} {\cal L}_D = \bar\psi(x) \, \left(i \gamma^\mu \partial_\mu - m \right) \, \psi(x) . \end{equation} Clearly, the free {\sc Dirac} Lagrangian is form invariant under global gauge transformations of the spinor wavefunctions, \begin{equation}\label{globaltrafo} \psi(x) \to \psi'(x) = e^{iq \alpha} \psi(x) , \qquad \bar\psi(x) \to \bar\psi'(x) = e^{-iq \alpha} \bar\psi(x) , \end{equation} with some arbitrary constant phase parameter $\alpha$ and charge $q$. The {\sc Noether} current corresponding to the transformations (\ref{globaltrafo}) is given by \begin{equation}\label{noethercurrent} \jmath^\mu = - q \, \bar\psi(x) \gamma^\mu \psi(x) . \end{equation} It satisfies the continuity equation, \begin{equation} \partial_\mu \jmath^\mu = 0 , \end{equation} which expresses the conservation of charge. In order to identify $q$ in (\ref{noethercurrent}) empirically with the elementary charge $e$, we have to couple the {\sc Dirac} particle -- perhaps an electron -- to the electromagnetic field. Thus, the free Lagrangian, which is an idealization anyway, must be replaced by some Lagrangian describing interaction. Miraculously, it turns out that this coupling can in fact be {\em derived} just by postulating the invariance of (\ref{dirac_L}) under local gauge transformations instead of the corresponding global ones (\ref{globaltrafo}). \begin{quote} {\em {\bf Gauge postulate.} The Lagrangian of a free matter field $\phi_i(x)$ should remain invariant under local gauge transformations $\phi_i(x) \to \phi'_i(x) = \phi'_i \left( \phi_i(x), \alpha_s(x) \right)$. } \end{quote} To see `how the miracle occurs' in the example, we consider the free {\sc Dirac} equation\footnote{This is the {\sc Euler-Lagrange} equation belonging to (\ref{dirac_L}).} \begin{equation}\label{dirac} (i \gamma^\mu \partial_\mu - m ) \, \psi(x) = 0 . \end{equation} Due to the gauge postulate we have to replace (\ref{globaltrafo}) by \begin{equation}\label{phasetrafo} \psi(x) \to \psi'(x) = e^{iq \alpha(x)} \psi(x) \end{equation} with a local, i.e. spacetime dependent, phase function $\alpha(x)$. Obviously (\ref{dirac}) is not invariant under this local gauge transformation. If we, however, identify \begin{equation} \label{phaseident} A_\mu(x) = - \partial_\mu \alpha(x) \end{equation} and thereby introduce a coupling field, which itself satisfies the local gauge transformations \begin{equation}\label{pottrafo} A_\mu(x) \to A'_\mu(x) = A_\mu(x) - \partial_\mu \alpha(x) , \end{equation} we may get a new interaction equation \begin{equation}\label{dirac_int} \left( i \gamma^\mu \partial_\mu - m \right) \psi(x) = q \ \gamma^\mu A_\mu(x) \ \psi(x) . \end{equation} This equation is indeed invariant under the combined transformations (\ref{phasetrafo}) and (\ref{pottrafo}). Formally it seems reasonable to identify $A_\mu$ with the electromagnetic potential\footnote{The experimental evidence for this maneuver is usually seen in the existence of the {\sc Aharonov-Bohm} effect, which should justify the crucial identification (\ref{phaseident}).}, since the construction of a field strength tensor (as the derivative of the potential) which is invariant under (\ref{pottrafo}) gives \begin{equation} F_{\mu\nu}(x) = \partial_\mu A_\nu(x) - \partial_\nu A_\mu(x) . \end{equation} This tensor satisfies the vacuum {\sc Maxwell} equations \begin{equation} \partial^\mu F_{\mu\nu}(x) = 0 \end{equation} and, as a {\sc Bianchi} identity, \begin{equation} \partial_{[\mu} F_{\nu\rho]}(x) = 0 . \end{equation} {\sc Maxwell}'s equations follow from the Lagrangian of the free electromagnetic field \begin{equation}\label{maxwell_L} {\cal L}_{EM} = - \frac{1}{4} F_{\mu\nu}(x) F^{\mu\nu}(x) . \end{equation} The coupling can explicitly be seen by introducing a {\em covariant derivative} \begin{equation}\label{ableitung} \partial_\mu \to D_\mu = \partial_\mu - iq A_\mu(x) \end{equation} and therefore \begin{equation} {\cal L}_{int} = - \jmath_\mu(x) A^\mu(x) . \end{equation} Since \begin{equation}\label{dirac_kovariant} (i \gamma^\mu D_\mu - m ) \ \psi(x) = 0 \end{equation} is equivalent to (\ref{dirac_int}), the coupling of matter and interaction fields can be derived in just one step via (\ref{ableitung}), thereby satisfying the gauge postulate by means of \begin{equation} {\cal L}_D \to {\cal L}' = {\cal L}_D + {\cal L}_{int} . \end{equation} \begin{quote} {\em {\bf Gauge principle.} The coupling of the {\sc Noether} current corresponding to the global gauge transformations of the Lagrangian of free matter fields can be introduced via replacing the usual derivative by the covariant derivative $\partial_\mu \to D_\mu$ corresponding to local gauge transformations. } \end{quote} Now, the merit of the concept of gauge field theories in modern physics becomes evident since the gauge principle provides a most successful and elegant `recipe' for introducing interaction, e.g. in our example deriving from the free theory (\ref{dirac_L}) via (\ref{phasetrafo}) the coupling structure of quantum electrodynamics,\footnote{In order to obtain full quantum electrodynamics the fields $\psi(x)$, $\bar\psi(x)$, and $A_\mu(x)$ have to be quantized.} \begin{equation}\label{QED_L} {\cal L}_{QED} = {\cal L}_D + {\cal L}_{EM} + {\cal L}_{int} . \end{equation} \section{Gauge transformations and gauge freedom} As we have seen, quantum electrodynamics can be understood as a gauge field theory proper with gauge group $U(1)$, since the gauge transformations occuring are global and local $U(1)$ transformations. Unfortunately, the usage of the terms `gauge theory' and `gauge transformations' is by no means uniform throughout the literature, which sometimes leads to conceptual confusions. In order to clarify the terminology I shall make some necessary distinctions: \begin{enumerate} \item Global gauge transformations, also called `gauge transformations of the first kind' ($GT1$), with corresponding gauge group $G1$. \item Local gauge transformations, also called `gauge transformations of the second kind' ($GT2$), with corresponding gauge group $G2$. \end{enumerate} On closer inspection of $GT2$ one should distinguish two kinds: \begin{itemize} \item[2a.] Matter field\footnote{This is the usual terminology, although there may exist fundamental particle fields with mass zero such as, perhaps, neutrinos -- `energy-matter field' is certainly the more precise term.} transformations, hereby called `type a' gauge transformations of the second kind ($GT2a$). \item[2b.] Gauge field\footnote{This is again the usual terminology, although `gauge potential' is more precise. The derivative gives the gauge field strength.} transformations, hereby called `type b' gauge transformations of the second kind ($GT2b$). \end{itemize} Regarding the example from the preceeding section the $GT2a$ are given by (\ref{phasetrafo}), whereas the $GT2b$ are given by (\ref{pottrafo}). Usually the $GT1$-$GT2$ distinction is made, seldom however $GT2a$-$GT2b$. One exception is for instance {\sc Wolfgang Pauli}, who in an early influential article concerning the gauge approach in relativistic field theories indicates the $GT2a$-$GT2b$ distinction, however calling it {\em ``... gauge transformations of the first ... and ... of the second type''} \cite[p.207]{pauli41}. I very much agree with {\sc Pauli} in regarding this as an important distinction, which is, as he points out, {\em ``... manifested through the fact that only expressions which are bilinear in $U$ and $U^*$ {\em [comparable with the matter fields $\psi$ and $\bar\psi$ in (\ref{dirac_L})] } are associated with physically measurable quantities ...''} and {\em ``... that, in principle, only gauge invariant quantities can be obtained by direct measurement''}.\footnote{{\sc Pauli} was always concerned by questions relating to the measurability of quantized fields. Already in \cite[p. 579]{pauli33} he noticed {\em ``... da\ss \ f\"ur das Photonfeld ... der Begriff der raum-zeitlich-lokalen Teilchendichte $W(\vec x,t)$ nicht sinnvoll existiert'' [ ... that for the photon field ... the notion of a particle density $W(\vec x,t)$ located in space-time has no meaningful existence]}. This is due to the fact that the four current for the photon field identically vanishes, $\jmath^\nu_{EM} = \partial_\mu \hat F^{\mu\nu} = 0$, which has considerable consequences for the interpretation of quantum fields. Since there is no local conservation law for the number of photons, the concept of a well-defined particle density is not in the same sense meaningful for the photon gauge field as it is for the {\sc Dirac} matter field \cite{lyre96a}.} Moreover, the structure of the $GT2b$ -- although they already appear in the free {\sc Maxwell} theory (\ref{maxwell_L}) -- is forced by the structure of the $GT2a$, thus, expressing the `miracle' of the gauge principle in other terms: the gauge field appears as an appendix of the matter field! In addition to the 3-fold distinction ($GT1$, $GT2a$, $GT2b$) concerning the usage of the notion of `gauge transformation', the term `gauge theory' should be clarified. It is already a common practice to call classical electrodynamics, i.e. the free {\sc Maxwell} theory, a gauge theory. This is due to the fact that (\ref{maxwell_L}) is form invariant under the $GT2b$ transformations (\ref{pottrafo}). But this is certainly a misleading terminology, since we better should refer to this invariance as a {\em gauge freedom} of the theory, whereas only the combined {\sc Dirac-Maxwell} theory, i.e. quantum electrodynamics (\ref{QED_L}), is to be considered as a true {\em gauge theory}. Note that the same argument also holds for the diffeomorphism invariance of our known spacetime theories. Whether general relativity, for instance, may nonetheless be considered as a gauge theory is another question and should not be confused with the obvious gauge freedom concerning the group of diffeomorphisms of the spacetime manifold. Thus, a theory comprising some gauge freedom is not yet a gauge theory, but only a theory incorporating the gauge principle. \section{The fiber bundle structure of gauge theories} \label{fb} I shall recall the definition of a fiber bundle $\langle \mathbbm E, \Man, \pi, \mathbbm F, G \rangle$ with bundle space $\mathbbm E$, base manifold $\Man$, projection map $\pi: \mathbbm E \to \Man$, fiber space $\mathbbm F$, and structure group $G$ -- compare e.g. \cite{nakahara90}. Fiber bundles can be considered as generalizations of direct product spaces, locally looking like $\Man\times \mathbbm F$. Thus, a local trivialisation is given by a diffeomorphic map $\phi_i: {\cal U}_i \times \mathbbm F \to \pi^{-1}({\cal U}_i)$ within some open set ${\cal U}_i \subset \Man$. In order to obtain the global bundle structure the local charts $\phi_i$ must be glued together via transition functions $t_{ij}(p) = \phi^{-1}_{i,p} \circ \phi_{j,p}$ with $\phi_{i,p}(f) \equiv \phi_i(p,f)$, $p \in \Man$, $f \in \mathbbm F$. A bundle section is a mapping $s: \Man \to \mathbbm E$ and can be considered as a generalization of a tangent vector field. With $\pi\left( s(p) \right) = p$ the section $s(p) \in \mathbbm F_p$ is local. A bundle is called trivial, if it admits a global section. In physics two classes of bundles play a central role. If the fiber is given by some $n$ dimensional linear vector space $\mathbbm V^n$ the bundle is called a vector bundle $\mathbbm E(\Man, \mathbbm V^n, GL(n,\mathbbm V) )$. For $\mathbbm V^n=\mathbbm R^n$ the general structure group is $G=GL(n,\mathbbm R)$. For principal bundles $\mathbbm P(\Man, G)$ the fiber is itself a {\sc Lie} group $\mathbbm F \equiv G$ with a natural action on the bundle from the right, $\mathbbm P \times G \to \mathbbm P$. To any principal bundle there naturally exists an associated vector bundle with the same structure group and transition functions. It is crucial for our considerations to understand that the geometrical framework of fiber bundles provides indeed a natural mathematical setting for the representation of physical gauge field theories. It turns out that any of the four fundamental interactions can be represented within the framework of a principal bundle. For instance, quantum electrodynamics in section \ref{gaugeprin} is to be considered as a $U(1)$ gauge field theory with a trivial bundle $\mathbbm P(\Mink, \mathbb S^1)$ -- the triviality of this bundle being due to the fact that {\sc Minkowski}an spacetime $\Mink$ is contractible. Within the fiber bundle language, gauge physical notions obey the following dictionary: A matter field is given as a section $s$ in the associated vector bundle ($\mathbbm E(\Mink, \mathbbm C, U(1) )$ in our example). In order to describe the gauge potential the concept of a bundle connection $\omega$ is needed: $\omega$ is a 1-form with values in the {\sc Lie} algebra $\lie{g}$ of the structure group $G$ which defines a unique decomposition $T\mathbbm P = V\mathbbm P \oplus H\mathbbm P$ of the tangent space $T\mathbbm P$ of the principal bundle $\mathbbm P$ into a `vertical' and a `horizontal' part. Note that the vertical subspace $V \mathbbm P$ is isomorphic to $\lie{g}$. Hence, the {\sc Dirac-Maxwell} gauge potential can be represented as $A = s^* \omega = A_\mu dx^\mu$, where $s^*$ is the pull back. The local gauge transformations (\ref{pottrafo}), which stem from the transition functions $t_{ij}$ above, may in general be written as $\omega' = g^{-1} \omega g + g^{-1} d g$ with $g \in G$. Thus, the structure group serves as the gauge group. It is important to note that, again, this terminology is not uniform. Due to the local gauge postulate the gauge group consists of spacetime-dependent group elements and is, thus, infinite dimensional. Therefore sometimes the gauge group is to be considered as a subgroup ${\cal G} \subset \AutP$ of the automorphism group of $\mathbbm P$. The `pure gauge transformations', which preserve the connection, are then given by the group ${\cal G}_o$ of just the vertical automorphisms. This is of particular interest when considering general relativity as a gauge theory. Here, the bundle structure is given by the orthonormal frame bundle $\LoM$ (sometimes called tetrad bundle since the frames are orthonormalized tetrads) with the homogeneous {\sc Lorentz} group as structure group. Trickily, as authors like {\sc Andrzej Trautman} have pointed out \cite{trautman80b}, in general relativity the `gauge group' $\cal G$ is simply isomorphic to the diffeomorphism group $\DiffM$ of the spacetime manifold $\Man$, whereas the group of `pure gauge transformations' shrinks to the identity ${\cal G}_o = id$. Although this is certainly an important characteristic of general relativity, I prefer calling the structure group the gauge group -- in contrast to $\cal G$ as the {\em group of local gauge transformations}. In gravitational gauge theories this refers to the central conceptual importance of tetradial reference frames, which are locally free to rotate and translate due to the structure group. Therefore sometimes general relativity can alternatively be considered as a translational gauge theory, since the tetrads represent the gauge potentials of the translation group isomorphic to $\Mink$. In this case the corresponding {\sc Noether} currents are given by the energy-momentum density current, a procedure which exactly mimics the way the gauge principle works.\footnote{These questions are closely related to the fact that in general relativity the fiber is soldered to the base manifold, which indicates the most important conceptual difference from quantum gauge field theories. Moreover, it seems quite natural to extend the structure group, which leads to generalized theories of gravitation -- most of them including torsion. Thus, orthodox general relativity fits into the gauge theoretic framework although this framework forces one to think about more general structures -- compare \cite{hehl_etal95} for an overview.} I shall now claim -- as an important feature of any gauge approach -- that the local gauge transformations $GT2$ only allow for an active interpretation. This can directly be seen from the gauge postulate. More precisely, this point of view holds inherently for the local gauge transformations of the matter fields $GT2a$, whereas the gauge fields turn out to be a consequence of this postulate (which is the very idea of the gauge principle). In general relativity, `matter fields' are represented in terms of reference frames. We therefore are forced to think about reference frames as the building blocks of gravitational gauge theories \cite{hehl94} -- even more so in encountering quantum gravity \cite{rovelli91b}. Concluding this, the following definition of gauge field theories can be given: \begin{quote} {\em {\bf Definition.} We call a {\bf\em gauge field theory} any theory being derived from the gauge principle and representing the geometry of a principal fiber bundle. The gauge group is given by the structure group of the bundle. } \end{quote} \section{Hole arguments till now} In 1913-1914, during his crucial years of developing general relativity, {\sc Einstein} claimed mistakenly that the new theory might not be generally covariant, i.e. form invariant under general coordinate transformations. He tried to convince himself by the so-called {\em ``Lochbetrachtung''}, i.e. by considering a hole in the matter energy distribution where $T_{\mu\nu}=0$, which would lead to the possibility of describing the metric $g_{\mu\nu}$ via, in modern terminology, different diffeomorphic tensors -- a possibility which for {\sc Einstein} seemed to contradict the law of causality. This is the first famous ``hole argument'' and its historical and philosophical background has been scrutinized in detail -- see e.g. \cite{norton84}, \cite{stachel89}. For the purpose of this paper only the structure of the argument will be of interest. It can be repeated in two steps: \begin{enumerate} \item {\em {\sc Leibniz} equivalence of diffeomorphic models of usual spacetime theories (general covariance):} In the relationalist's view, diffeomorphic models of any spacetime theory making use of the concept of smooth manifolds to represent spacetime are equivalent with regard to any observation, i.e. they represent one and the same physical situation. \item {\em Failure of determinism:} The substantivalist's assumption of an existence of spacetime points independent from the matter content of spacetime leads to an indeterminism by considering different diffeomorphic models of a theory whose predictions cannot be used to make out any empirical distinction between the `different' models. Thus, indeterminism arises due to the substantivalist's denying of {\sc Leibniz} equivalence. \end{enumerate} In 1987, {\sc John Earman} and {\sc John Norton} presented a new hole argument following this 2-step structure, which is valid for the whole class of spacetime theories containing diffeomorphic models \cite{earman+norton87}. In general, such a model is given by a tupel $\langle \Man, O_1, ... O_n \rangle$ with a spacetime manifold $\Man$ and quantities $O_1, ... O_n$ denoting certain geometric objects. Thus, a model of general relativity is given by $M=\langle \Man, g_{\mu\nu}, T_{\mu\nu} \rangle$ with metric $g_{\mu\nu}$ and energy-momentum tensor $T_{\mu\nu}$. The authors claim a {\em gauge theorem} which reads as follows: \begin{quote} ``Gauge Theorem {\em (General covariance): If $\langle \Man, O_1, ... O_n \rangle$ is a model of a local spacetime theory and $h$ is a diffeomorphism from $\Man$ to $\Man$, then the carried along tuple $\langle \Man, h^* O_1, ... h^* O_n \rangle$ is also a model of the theory.''}\cite[p.~520]{earman+norton87} \end{quote} Recall that any diffeomorphism $f: \Man \to \Man$ induces a map (carry along) $f^*: \mathbbm V_p \to \mathbbm V_{f(p)}$ at points $p \in \Man$, which means that any tensor $T$ given in the coordinates $\left\{ x^i \right\}$ of $\mathbbm V_p$ is also given as $f^* T$ in new coordinates $\left\{ y^i \right\}$ of $\mathbbm V_{f(p)}$. Hence, the model $f^*M=\langle \Man, f^* g_{\mu\nu}, f^*T_{\mu\nu} \rangle$ is {\sc Leibniz} equivalent to $M$, i.e. $M$ and $f^*M$ are empirically indistinguishable. This is the first part of the argument. For the second part {\sc Earman} und {\sc Norton} chose a special {\em ``hole diffeomorphism h''} with \begin{equation} \label{hole} h=id, \quad t \le t_o, \qquad \mbox{and} \qquad h \ne id, \quad t>t_o , \end{equation} which, of course, obeys usual smoothness and differentiability conditions at $t_o$. Hence, we have $M = h^*M$ for times $t \le t_o$, whereas $M \ne h^*M$ for $t>t_o$. Since the spacetime substantivalist must claim that at $t_o$ the world splits into two physically distinct models $M$ and $h^*M$ -- although the theory cannot predict any empirical difference --, for him a radical inherent indeterminism arises. It is important to note that {\sc Earman} and {\sc Norton} for the first step of the argument consider active point transformations instead of mere passive coordinate transformations. They indicate this by the term ``gauge theorem'', which they explain by quoting {\sc Robert Wald}: {\em ``... the diffeomorphisms comprise the gauge freedom of any theory formulated in terms of tensor fields on a spacetime manifold''} \cite[p.~438]{wald84}. Thus, by converting the notion of general covariance into an active language of point transformations, the hole argument becomes vivid. After all, we are confronted with the following alternatives in order to escape the hole argument: either to give up spacetime substantivalism, or to accept indeterminism as a consequence of any generally covariant spacetime theory. But the latter way out, {\sc Earman} und {\sc Norton} close, seems to be {\em ``... far too heavy a price to pay for saving a doctrine that adds nothing empirically to spacetime theories''} \cite[p.~524]{earman+norton87}. \section{The generalized hole argument} The above manifold hole argument rules out spacetime substantivalism for orthodox spacetime theories including general relativity. As shown in section \ref{fb}, in gauge field theories one naturally has to take into account the enlarged geometrical arena of the underlying fiber bundles to represent matter and gauge fields. The question arises, which kind of status is appropriate for to the geometry of fiber bundles. In particular: Does a relationalistic or a substantivalistic point of view hold? In order to rule out the latter one, I argue in the following that there exists a straightforward extension of the spacetime manifold hole argument to a generalized bundle space hole argument. One first of all should ask whether there exist reasons at all for believing in fiber bundle substantivalism. Since the spacetime hole argument already rules out manifold substan\-ti\-va\-lism, the fiber bundle substantivalist will claim the independent existence of fiber spaces as internal geometrical spaces in which matter and gauge fields `live'. First, this is the decisive argument for the general need of fiber bundles: Fields do not live in spacetime itself; rather, they live in state spaces defined on spacetime.\footnote{This statement surely holds for quantum gauge field theories in our standard model. Whether this is even true for the connection forms, i.e. the gauge fields, of general relativity, is of course a matter of a more detailed analysis and interpretation of the theory's underlying gauge structure. In the light of the few remarks at the end of section \ref{fb}, I like to assume this to be the case: Gauge fields of gravity -- as well as their derived properties such as curvature -- live primarily in the fibers. They are constituted by actual transformations of local reference frames. Speaking about the curvature of the manifold, though, appears from the gauge theoretic point of view as a conventional maneuver due to the fact that fiber and base space are soldered. Surely, a detailed discussion of this topic remains to be a further task.} As indicated in section \ref{fb}, at each spacetime point we need two additional spaces constituting the geometrical arena of our world: A group space constituting the state space of gauge fields, which is provided by some principal bundle $\mathbbm P$, and a vector space for matter fields, provided by the associated vector bundle $\mathbbm E$. Now, the fiber bundle substantivalist will consider bundle, or at least fiber-space points (if he has already accepted the spacetime hole argument) as individuated substances. Analogous to the usual spacetime substantivalist's emphasis on the existence of vacuum solutions in general relativity, the fiber space substantivalist could point out the activity and effectiveness of the vacuum in quantum gauge field theories. Indeed, {\sc Yuval Ne{'e}man} claims, that {\em ``... the vacuum is ... the arena for the nonlinear interaction of the gauge fields. As a result, spacetime is a physical entity -- as a set of fields -- at the classical level already.''} Therefore, {\em ``... physics selects the realist or substantivist view, and contradict[s] the tenets of relationalism or conventionalism, with respect to spacetime ...''} \cite{neeman95}. Curiously enough, {\sc Ne{'e}man} applies his argument to spacetime alone -- obviously ignoring the spacetime hole argument --, although, if at all, his argument should a fortiori hold for fiber spaces, too. I shall now confront these points of view with the relationalist's arguments. This will be done in the same two steps as for the spacetime hole argument. However, I like to argue that one does not necessarily need a second step. At least for the case of principal bundles the first step of the argument is already sufficient. Thus, interestingly, one does not need a proper hole argument in this case. \subsection{Generalized hole argument: First step} Consider the usual gauge freedom arising in any gauge theory $T$. Thus, $T$ admits gauge transformations $GT1: T \to T'$ and $GT2b: T(x) \to T'(x)$ such that $T$ and $T'$ resp. $T(x)$ and $T'(x)$ are {\sc Leibniz} equivalent. In other words, only gauge invariant quantities are observable. In order to make empirical use of $T$, it is necessary to fix a gauge. This is evidently just a conventional operation, such as introducing coordinates. The gauge by its own nature has no significant physical meaning. Surely, the substantivalist will not deny this, but he will nevertheless insist on the ontological individuality of points in spaces in which the gauge is applied. As demonstrated, the {\sc Earman-Norton} hole argument makes a decisive use of an active interpretation of the considered transformations as point transformations instead of passive coordinate transformations. In order to take over the argument I shall refer to $GT2b$ as point transformations in the group manifold of $G$. This reflects the very nature of the fiber spaces in the framework of a principal bundle $\mathbbm P$: The right action of $G$ on $\mathbbm P$ leads to the fibration of the bundle, i.e. $G$-orbits (fibers) are equivalence classes of physically indistinguishable states. Here, a crucial point arises: Since the fiber space is the group itself, its points have {\em per definition} no significant physical meaning as entities per se. One can see this by recalling the idea of thinking about {\sc Lie} groups in terms of their parameter manifolds. This leads to a most natural representation: the {\sc Lie} group as its own homogeneous space, thus, the group acting transitively on itself. From this point of view we are forced to primarily interpret the abstract group in terms of its algebraic rather than its analytic structure, which seems to be the natural way of an application of {\sc Lie} groups within the framework of principal bundles. Therefore -- since gauge fields just take values in the group -- merely a relationalist's interpretation of the group's natural homogeneous representation space holds: no points are distinguished, and, moreover, no group space point has any physical significance whatsoever. It is indeed the key idea of gauge theories that only relations of gauge transformations within different fibers, given by the connection forms, have any empirical meaning. Beside the structure of principal bundles $\mathbbm P$, gauge theories make use of their associated vector bundles $\mathbbm E$ as well. How to proceed with their fiber spaces? Clearly, it is one and the same abstract group $G$ which constitutes the fibers of $\mathbbm P$ and acts on the vector space fibers of $\mathbbm E$. The matter field `lives' in the latter ones and is thereby just a representation of $G$ in some vector space.\footnote{E.g., each component $\psi^i$ of the {\sc Dirac} bispinor $\psi$ gives a fundamental representation of $U(1)$ in $\mathbbm C$.} Matter fields transform according to local gauge transformations $GT2a$. In section \ref{fb}, I gave an interpretation of $GT2a$ as inherently active transformations, namely, a local change of a matter field at some point $p$ compared to a different point $p'$ in spacetime changes the physical situation, i.e. constitutes an interaction represented as a gauge field. Changing the physical situation can be understood as the general meaning of transformations considered to be active. Note however, that this does not necessarily refer to active {\em point} transformations (of spaces whatsoever). Actually, in stressing local gauge transformations being actively interpreted, we are by no means forced to consider them as point transformations -- they rather represent active changes of general state space reference frames. The vector bundle substantivalist, however, will consider local gauge transformations $GT2a$ as active point transformations in the vector space fibers. The relationalist's arguments must then prove this point of view to be untenable. However, as we will see now, the mere representation of $G$ in the vector space fibers of $\mathbbm E$ does not allow for the same argument to rule out vector bundle substantivalism as for the group manifold fibers of $\mathbbm P$. The gauge fixing mentioned above acts in the vector space as distinguishing a certain basis. Again, this is a pure conventional maneuver: only local changes of state space reference frames have a physical meaning. But, the substantivalist may still claim that vector space points are entities per se, since gauge fixing is related to a mere passive operation such as choosing coordinates. Here, the relationalist has to accept that representing the bundle's structure group in a vector space is not reason enough to derive a pure relational status of such a space, simply because the group theoretic argument as in the principal bundle case above does not apply. Thus, the first step of the generalized argument is only sufficient to rule out principal fiber bundle substantivalism, since these fiber spaces are, unlike base manifolds or any kind of vector spaces, group manifolds, i.e. spaces in which the structure group is homogeneously represented on itself. Regardless of the question about an approriate active interpretation of local gauge transformations, I see no way to individuate the points of such kinds of spaces. With regard to vector bundle fiber spaces, our mere mathematical tools, however, do still allow a substantivalist's viewpoint - even if this position turns out to be absurdly extreme. We can make this clear by emphasizing an inherent fundamental conventionalism in any gauge physics: we must fix a gauge in order to make empirical use of a gauge theory. This clearly indicates (but does not prove) that fiber space points are physically indistinguishable and non-individuated {\em because} of our freedom of choosing a particular gauge. \subsection{Generalized hole argument: Second step} I shall now confront `hard core substantivalists', who are still not convinced, with the proper version of the generalized hole argument. For this purpose, I will consider bundle isomorphisms instead of base manifold diffeomorphisms. Recall the following commutative diagram: \begin{equation} \begin{array}{rcccl} & \mathbbm E & \stackrel{\phi}{\longrightarrow} & \mathbbm E' & \\ {\scriptstyle \pi} & \downarrow & & \downarrow & {\scriptstyle \pi'} \\ & \Man & \stackrel{f}{\longrightarrow} & \Man' \end{array} \end{equation} If $\phi$ is a diffeomorphism, we may call it a bundle isomorphism. Per definition, bundle isomorphisms preserve the fiber structure of the bundle. In particular, $\phi: \mathbbm E \to \mathbbm E$ is a bundle automorphism. It can be read from the diagram that any bundle isomorphism uniquely induces a manifold diffeomorphism $f: \Man \to \Man'$. I shall now choose an appropriate ``hole isomorphism $\tau$''\footnote{There is no Greek counterpart of the letter `$h$' -- the reader may guess why `$\tau$' is chosen instead ...}. To begin with, one simply might use a bundle isomorphism which induces the hole manifold diffeomorphism (\ref{hole}) -- this already would be sufficient. But one may even think about a most general hole isomorphism \begin{equation} \tau=id, \quad t \le t_o, \qquad \mbox{and} \qquad \tau \ne id, \quad t>t_o. \end{equation} In this way we are able to perfectly take over the second step of the hole argument: Since for the fiber space substantivalist the action of $\tau$ changes the `real' arrangement of bundle space points, i.e. the physical situation, the world splits again into different models, thus, leading to indeterminism. Note that this kind of indeterminism has nothing to do with the type of indeterminism arising in quantum theories (and, thus, in quantum gauge field theories). The {\sc Dirac-Maxwell} or, in general, {\sc Yang-Mills} field equations, which govern the temporal development of the fields are strict deterministic field equations. Therefore, the existence of symmetry properties of the fields such as bundle morphisms are clearly not related to indeterminism arising in the quantum measurement process. Hence, it should have become clear from the above arguments that there is no possibility left for the substantivalist to hold his position, since the proper use of bundle isomorphisms in the generalized hole argument rules out fiber bundle substantivalism in the same manner as base manifold diffeomorphisms rule out manifold substantivalism. Moreover, since it can be argued that the second part of the argument is not necessary at least in the principal bundle case, the substantivalist's possible escape into indeterminism is even more eroded. Thus, one ends up with a clear result: Fiber bundles refer to a relationalistic interpretation. \section{The meaning of gauging?} Once again, the idea of an active interpretation of local gauge transformations refers to the active relational change of reference frames of the matter fields. This considerably changes the physical situation, whereas the idea of actively shifting fiber space points remains without any empirical meaning because of these points being genuinly a representation of a group. Due to the gauge principle gauge fields appear to be a consequence and, thus, a mere appendix of the matter fields. This is another way of arguing that the notion of a matter free spacetime is without any empirical meaning. Since we must regard the gauge principle as a tremendous successful heuristic principle in modern theoretical physics, the more so we should be puzzled with the unsolved philosophical question concerning the meaning of this principle. Until today it still remains a pure miracle why the postulate of local gauge transformations, i.e. replacing the transformation parameters $\alpha_s \to \alpha_s(x)$, leads to the coupling of matter and interaction fields. It seems quite clear that hand waving arguments such as ``field physics has to be local, therefore the transformations must be local'', which one finds throughout the textbook literature, are philosophically by no means satisfying. Maybe, the curious interplay between global and local considerations in the gauge approach gives us a hint for considering new ideas of spacetime -- not referring to it as being primarily a differentiable manifold \cite{lyre98}. But these questions touch the deep conceptual roots of physics in general. At this stage, the real puzzle begins and therefore a lot of work needs to be done by physicists as well as philosophers of physics to find the true meaning of gauging. \vspace{1cm} \subsection*{Acknowledgements} I would like to thank the people at the Center for Philosophy of Science of the University of Pittsburgh for their kind hospitality and support during my stay as a Visiting Fellow in the 1998-99 academic year. Moreover, I thank {\sc Michael Drieschner}, {\sc Tim Eynck}, {\sc Yair Guttmann}, {\sc Allen Janis}, and {\sc John Norton} for helpful remarks. This paper has been made possible by a scholarship from the Alexander~von~Humboldt-Foundation, Bonn. \newpage
\section{Introduction} Recent high precision experiments require, on the side of the theory, high-precision calculations resulting in the evaluation of higher loop dia\-grams in the Standard Model (SM). For specific processes thousands of multiloop Feynman dia\-grams do contribute, and it turns out to be impossible to perform these calculations by hand. This makes the request for automation a high-priority task. Several different packages have been developed with different areas of applicability (see a good review \cite{Steinh1}). For example, FEYNARTS / FEYNCALC \cite{FeynmArts} are MATHEMA\-TICA packages convenient for various aspects of the calculation of radiative corrections in the SM. There are several FORM packages for evaluating multiloop diagrams, like MINCER \cite{MINCER}, and a package \cite{leo96} for the calculation of 3-loop bubble integrals with one non-zero mass. Other packages for automation are GRACE \cite{GRACE} and COMPHEP \cite{CompHep}, which partially perform full calculations, from the process definition to the cross-section values. A somewhat different approach is pursued by XLOOPS \cite{XLoops}. A graphical user interface makes XLOOPS an `easy-to-handle' program package, but is mainly aimed to the evaluation of single diagrams. To deal with thousands of diagrams, it is necessary to use special techniques like databases and special controlling programs. In \cite{Vermaseren} for evalua\-ting more than 11000 diagrams the special database-like program MINOS was developed. It calls the relevant FORM programs, waits until they fi\-nished, picks up their results and repeats the process without any human interference. The package GEFICOM \cite{GEFICOM}, developed for computation of higher order processes involving a large number of diagrams, is based on cooperative usage of several software tools such as Mathematica, FORM, Fortran, etc. It seems impossible to develop an universal package, which will be efficient for all tasks. It appears absolutely necessary that various groups produce their own solutions of handling the problem of automation: various ways will be of different efficiency, have different domains of applicability, and last but not least, should eventually allow for completely independent checks of the final results. This point of view motivated us to seek our own way of automatic evaluation of Feynman diagrams. Our first step was dedicated to the automation of the muons two-loop anomalous magnetic moment (AMM) ${\frac{1}{2}(g-2)}_{\mu}$. For this purpose the package TLAMM was developed \cite{TLAMM}. The algorithm is implemented as a FORM-based program package. For generating and automatically evaluating any number of two-loop self-energy diagrams, a special C-program has been written. This program creates the initial FORM-expression for every diagram generated by QGRAF \cite{QGRAF}, executes the corresponding subroutines and sums up the various contributions. In the SM 1832 two-loop diagrams contribute in this case. The calculation of the bare diagrams is finished. Our aim is to create some universal software tool for piloting the process of generating the source code in multi-loop order for analytical or numerical evaluations and to keep the control of the process in general. Based on this instrument, we can attempt to build a complete package performing the computation of any given process in the framework of any concrete model. \section{Brief description} For the project called DIANA (DIagram ANAlyser) for the evalua\-tion of Feynman diagrams we have elaborated a special text manipulating language (TM). The TM language is a very simple TeX-like language for creating source code and organizing the interactive dialog. The program reads QGRAF output. For each diagram it performs the TM-program, producing input for further evaluation of the diagram. Thus the program: Reads QGRAF output and for each diagram it: \begin{enumerate} \item Determines the topology, looking for it in the table of all known topologies and distributes momenta. If we do not yet know all needed topologies, we may use the program to determine missing topologies that occur in the process. \item Creates an internal representation of the diagram in terms of vertices and propagators. \item Executes the TM-program to insert explicit expressions for the vertices, propagators etc. The TM-program produces input text for FORM ( or some other language), and executes the latter (optionally). \end{enumerate} Using the TM language, advanced users can develop further extensions, e.g. including FORTRAN, to create a postscript file for the picture of the current diagram, etc.\\ \begin{figure}[ht] \centerline{\epsfysize=130mm \epsfbox{dianam3.eps}} \caption{\label{dianam} Typical flowchart of Feynman diagram evaluations by DIANA. } \end{figure} The program operates as follows: first of all, it reads its configuration file, which may be produced manually or by DIANA as well. This file contains: \begin{enumerate} \item The information about various settings (file names, numbers of external particles, definition of key words, etc.) \item Momenta distribution for each topology. \item Description of the model (i.e., all particles, propagators and vertices). \item TM-program. \end{enumerate} The TM-program is part of the configuration file. It starts with the directive \small\begin{verbatim} \begin translate \end{verbatim}\normalsize Then the program starts to read QGRAF output. For each diagram it determines the topology, assigns indices and creates the textual representation of the diagram corresponding to the Feynman integrand. All defined data (masses of particles, momenta on each lines, etc.) are stored in internal tables, and may be called by TM-program operators. At this point DIANA performs the TM-program. After that it starts to work with the next diagram. When all diagrams are processed, the program may perform the TM-program a last time (optionally). This may be used to do some final operations like summing up the results. Use of the TM language makes DIANA very flexible. It is easy to work out various algorithms of diagram evaluation by specifying settings in the configuration file or even by a TM program. The typical flowchart may look like in Fig. \ref{dianam}. There is a possibility to use DIANA to perform the TM-program only, without reading QGRAF output. If one specifies in the configuration file \small\begin{verbatim} only interpret \end{verbatim}\normalsize then DIANA will not try to read QGRAF output, but immediately enters the TM - program. DIANA contains a powerful preprocessor. The user can create macros to hide complicated constructions. Similar as LaTeX provides the possibility for non-specialists to typeset high-quality texts using the TeX language, these macros permit DIANA to work at very high level. The user can specify the model and the process, and DIANA will generate all necessary files. \section{Topologies and momenta} \label{topologies} \begin{figure}[ht] \centerline{\epsfysize=40mm \epsfbox{topol1.eps}} \caption{\label{topol1} Topology (-2,2)(-1,1)(5,1)(1,3)(3,4)(4,2)(2,5)(5,6)(3,6)(4,6). } \end{figure} Topologies are represented in terms of ordered pairs of numbers like (fromvertex, tovertex) (see Fig.~\ref{topol1}). All external legs have negative numbers. These are supposed to begin with -1. First one attaches all the numbers for the ingoing particles, and then for the outgoing ones. External legs must be connected with vertices of smallest possible identifying number. The number of an internal line corresponds to its position in the chain of pairs and the direction from the first to the second number in the pair: $$ \begin{array}{ccccccccc} \mbox{direction:}&5\to 1&1\to 3&3\to 4&4\to 2&2\to 5&5\to 6&3\to 6&4\to 6\\ (-2,2)(-1,1)&(5,1)&(1,3)&(3,4)&(4,2)&(2,5)&(5,6)&(3,6)&(4,6)\\ \mbox{number:}&1&2&3&4&5&6&7&8 \end{array} $$ Knowing thus the topology we can assign momenta. Their distribution according to Fig. \ref{topol1}, e.g., is added like \small\begin{verbatim} topology A = (-2,2)(-1,1)(5,1)(1,3)(3,4)(4,2)(2,5)(5,6)(3,6)(4,6): p1,p2,p3,p4,p5,p6,p7,p8; \end{verbatim}\normalsize This fixes directions and values of all momenta on internal lines. External lines must be known from the process definition. By convention the external lines must start with the highest negative number. All topologies have to be described in the configuration file. DIANA stores topologies in an internal table in some standard form. After reading a new diagram DIANA defines its topology, reduces the topology to the standard form and searches for the topology in the table. If DIANA fails to find the topology, it will not produce the Feynman integrand for the diagram. \section{TM language} Similar to the TeX language, all lines without special escape - characters (``$\backslash$'') are simply typed to the output file. So, to type ``Hello, world!'' in the file ``hello'' we may write down the following program: \small\begin{verbatim} \program \setout(hello) Hello, world! \end{verbatim}\normalsize The asterisk in the beginning of line is a comment: \small\begin{verbatim} * This is the comment! \end{verbatim}\normalsize Each word, the first character of which is the escape character, will be considered as a command. This feature makes this language very easy-to-use. In appendix A there is an example of a simple TM-program and the result of its performance. DIANA's preprocessor is run on the TM program before actual compilation. It permits the user to do textual substitutions (with parameters), to perform conditional compilation, etc. For example, the following preprocessor directive \small\begin{verbatim} \DEF(macroname) . . . \ENDDEF \end{verbatim}\normalsize sets the macrodefinition. After this directive you can just write \small\begin{verbatim} \macroname() \end{verbatim}\normalsize to invoke the macro. You can use macros with arguments: \small\begin{verbatim} \macroname(a,b,c) \end{verbatim}\normalsize In the macro body the arguments are available by the directive \small\verb|\#(n)|, where ``n'' is the position of the argument. The argument ``0'' is the macro name, all extra arguments are empty strings. Example: \small\begin{verbatim} only interpret \begin translate \DEF(EXAMPLE) 0=\#(0),1=\#(1),2=\#(2),3=\#(3)\ENDDEF \program \offblanklines \setout() \EXAMPLE(a,b) \end translate \end{verbatim}\normalsize The result is: \small\begin{verbatim} 0=EXAMPLE,1=a,2=b,3= \end{verbatim}\normalsize Using macros we are able to create a high level macrolanguage similar to LaTeX, which is the set of macrodefinitions under TeX. The basic idea is again similar to LaTeX, namely the use of {\it styles}. Instead of complicated TM programming one can use proper styles. The style file containing all macrodefinitions is just included in the beginning of the TM program. We have several such styles oriented to the use of FORM. \section{Generating configuration files} The style ``create.tml'' has the purpose to create the TM program. The user has to prepare a file ``process.cnf'', which contains the model and process specifications. Reading this file, DIANA will generate all necessary other files. A typical example of such a file is the following. \small\begin{verbatim} SET _processname = "zbb" SET _qgrafname = "qgraf" SET _syspath = "/home/U.Ser/diana/tml/" system path = "/home/U.Ser/diana/tml/" only interpret \openlanguage(create.tml) \Begin(program) \Begin(model,gwsmassless.model) \End(model) \Begin(process) ingoing Z(mu;p1); outgoing b(;k1),B(;k2); loops = 2; \End(process) \Begin(qgrafoptions) options=onepi,nosnail; \End(qgrafoptions) \Begin(tmlprogram,form.prg) \End(tmlprogram) \indices(mu,mu1,mu2,mu3,mu4,mu5,mu6,mu7, mu8,mu9,mu10,mu11,mu12,mu13,mu14,mu15,mu16,mu17,mu18,mu19,mu20,mu21,mu22,mu23) \vectors(p1,p2,p3,p4,p5,p6,p,q,q1,q2,k1,k2) \End(program) \end{verbatim}\normalsize This file contains all necessary information to create files for DIANA to proceed the process $ Z \to b \bar b $ in the framework of the SM. In the following we explain details of the above file. \bigskip Directives of the type \small\verb|SET var = "val"|\normalsize{} are used to set predefined preprocessor variables. Thus, the directive \small\begin{verbatim} SET _processname = "zbb" \end{verbatim}\normalsize just defines the name of the process. It will be used as filename extension for all created files. \small\begin{verbatim} SET _qgrafname = "qgraf" \end{verbatim}\normalsize defines the command to invoke QGRAF. The directives \small\begin{verbatim} SET _syspath = "/home/U.Ser/diana/tml/" system path = "/home/U.Ser/diana/tml/" \end{verbatim}\normalsize are used to define the path to files containing TM macros and subroutines. The value of the preprocessor variable \small\verb|_syspath|\normalsize{} will be used by DIANA to define the line \small\begin{verbatim} system path = "/home/U.Ser/diana/tml/" \end{verbatim}\normalsize in the created file ``settings.zbb''. In principle in ``process.cnf'' and ``settings.zbb'' these two paths may be different. \bigskip The directive \small\begin{verbatim} only interpret \end{verbatim}\normalsize tells DIANA not to read QGRAF output but just to perform the TM program. \bigskip The directive: \small\begin{verbatim} \openlanguage(create.tml) \end{verbatim}\normalsize is equivalent to the two directives \small\begin{verbatim} \begin translate \include(create.tml) \end{verbatim}\normalsize The first one starts the TM language and the file \small\verb|create.tml|\normalsize{} contains all macro definitions and subroutines. In particular, it contains the definition of the macros \small\verb|\Begin()|\normalsize{} and \small\verb|\End()|. These macros form an environment depending on the first argument of the macro \small\verb|\Begin()|. \bigskip The Body of the TM program is supposed to be placed between \small\begin{verbatim} \Begin(program) . . . \End(program) \end{verbatim}\normalsize Text after \small\verb|\End(program)|\normalsize{} will be ignored. \bigskip The environment \small\begin{verbatim} \Begin(model,gwsmassless.model) \End(model) \end{verbatim}\normalsize defines the model description. Instead of explicit definition, in this example the file ``gwsmassless.model'' is used. This is a file containing a simplified version of the SM. It should be placed in the directory /home/U.Ser/diana/tml/ in order that DIANA can find it in ``process.cnf''. \bigskip The process definition is clear from \small\begin{verbatim} \Begin(process) ingoing Z(mu;p1); outgoing b(;k1),B(;k2); loops = 2; \End(process) \end{verbatim}\normalsize \bigskip The following environment defines the options passed to QGRAF \cite{QGRAF}: \small\begin{verbatim} \Begin(qgrafoptions) options=onepi,nosnail; \End(qgrafoptions) \end{verbatim}\normalsize \bigskip The TM program for evaluating Feynman diagrams must be placed in the environment \small\begin{verbatim} \Begin(tmlprogram) \End(tmlprogram) \end{verbatim}\normalsize Instead one may specify the file containing the TM program as second argument: \small\begin{verbatim} \Begin(tmlprogram,form.prg) \end{verbatim}\normalsize Here ``form.prg'' is the name of the file containing the standard TM program for evaluation of Feynman diagrams by means of FORM. \bigskip The macros \small\verb|\indices()|\normalsize{} and \small\verb|\vectors()|\normalsize{} specify used indices and vectors. \section{The structure of the generated files} Suppose the executable files ``diana'' and ``qgraf'' are available from the system path and the files ``create.tml'', ``gwsmassless.model'', ``form.prg'' and ``process.cnf'' are situated in the directory ``/home/U.Ser/diana/tml/'' or in the current directory. The user starts DIANA by means of the command \small\begin{verbatim} diana -c process.cnf \end{verbatim}\normalsize DIANA then generates several temporary files, invokes QGRAF, defines all topologies of the process and distributes momenta as follows: the first line in each topology will carry momentum p1, the second p2, etc. Three new files appear in the current directory: \mbox{``config.zbb''}, ``qlist.zbb'' and ``settings.zbb''. The file ``config.zbb'' contains the TM program for the evaluation of the Feynman diagrams of the specific process under consideration. ``settings.zbb'' contains various optional settings, topologies and momenta description, the model and the process definition. It can be edited e.g. for the purpose of introducing proper integration momenta for the various topologies by replacing the pi's. The file ``qlist.zbb'' is the QGRAF output file. To start the calculation of the Feynman diagrams the user finally enters the command \small\begin{verbatim} diana -c config.zbb \end{verbatim}\normalsize For the above example $ Z \to b \bar b $ the topologies in the file ``settings.zbb'' are: \small\begin{verbatim} topology top1_= (-3,3)(-2,2)(-1,1)(1,2)(1,3)(1,4)(2,4)(3,4): p1,p2,p3,p4,p5; topology top2_= (-3,3)(-2,2)(-1,1)(1,3)(1,4)(1,4)(2,3)(2,4): p1,p2,p3,p4,p5; topology top3_= (-3,3)(-2,2)(-1,1)(1,2)(1,4)(1,4)(2,3)(3,4): p1,p2,p3,p4,p5; topology top4_= (-3,3)(-2,2)(-1,1)(1,4)(1,4)(2,3)(2,4)(3,4): p1,p2,p3,p4,p5; topology top5_= (-3,3)(-2,2)(-1,1)(1,4)(1,5)(2,4)(2,5)(3,4)(3,5): p1,p2,p3,p4,p5,p6; topology top6_= (-3,3)(-2,2)(-1,1)(1,2)(1,4)(2,5)(3,4)(3,5)(4,5): p1,p2,p3,p4,p5,p6; topology top7_= (-3,3)(-2,2)(-1,1)(1,3)(1,4)(2,4)(2,5)(3,5)(4,5): p1,p2,p3,p4,p5,p6; topology top8_= (-3,3)(-2,2)(-1,1)(1,4)(1,5)(2,3)(2,4)(3,5)(4,5): p1,p2,p3,p4,p5,p6; topology top9_= (-3,3)(-2,2)(-1,1)(1,2)(1,3)(2,4)(3,5)(4,5)(4,5): p1,p2,p3,p4,p5,p6; topology top10_= (-3,3)(-2,2)(-1,1)(1,2)(1,4)(2,3)(3,5)(4,5)(4,5): p1,p2,p3,p4,p5,p6; topology top11_= (-3,3)(-2,2)(-1,1)(1,3)(1,4)(2,3)(2,5)(4,5)(4,5): p1,p2,p3,p4,p5,p6; \end{verbatim}\normalsize The user must edit this file to set the desired integration momenta and the internal numeration of lines according to the rules described in Sect. \ref{topologies}. In the near future the pictorial representation of the topologies will be implemented. \bigskip As next we give a detailed description of the file ``config.zbb''. It contains the directive to include the file ``settings.zbb'', several settings for various options and a skeleton of the FORM program in terms of the TM language. The following is a listing of the file. The whole TM-program is divided into different ``sections''. Each section will be performed only under proper conditions. \begin{quote} \small\begin{verbatim} * This file is automatically generated by DIANA for the process zbb. \include(settings.zbb) *Remove the following to avoid the generation of the protocol file: log file = log.zbb * Remove the following line to avoid debug information: debug on extra call \openlanguage(specmode.tml) \Begin(program,routines.rtn) \section(regular) \Begin(initialization) \export(callform,\ask(\(Call form?(Y/N)))) \End(initialization) \Begin(output,\askfilename()) **** d\counter() **** (diagram \currentdiagramnumber()) \blankline() * Set here your defines! \Begin(foreach,i,1,\numberofinternallines()) #define mm\i() "\mass(\i())" \blankline() \End(foreach) #define LINE "\numberofinternallines()" #define FERMIONLINE "\maxfcount()" #define TOPOLOGY "\topologyid()" \blankline() functions \functions(); commuting \commuting(); l Rq =\integrand(); \blankline() * Here should be your FORM program! drop Rq; g dia\counter()=Rq; .store save dia\counter().sto; .end \End(output) \Begin(special,"\import(callform)"eq"Y") \execute(\(form -l )!.!) \End(special) \section(epilog) \Begin(output) #do j = 1,\counter() load dia'j'.sto; .store #enddo g Rq = #do j = 1,\counter() + dia'j' #enddo ; .sort print; .end \End(output) \Begin(special,"\import(callform)"eq"Y") \execute(\(form -l )!.!) \End(special) \End(program) \end{verbatim}\normalsize \end{quote} The setting \small\verb|log file = log.zbb|\normalsize{} induces the creation of the protocol file ``log.zbb''. \bigskip If the user sets \small\verb|debug on|, DIANA will produce a detailed record of the translation process. Some information necessary for possible run-time diagnostics will be stored as well. \bigskip The setting \small\verb|extra call|\normalsize{} forces DIANA to perform the TM-program once again after all diagrams are processed. In this case only the section \small\verb|epilog|\normalsize{} will be called (see below). In the next command the TM-program uses the macro - style ``specmode.tml'': \small\begin{verbatim} \openlanguage(specmode.tml) \end{verbatim}\normalsize This style defines the ``sections'' mechanism: a TM-program may consist of several ``sections'', each of them being performed only under certain conditions. So the generic structure of the TM-program looks like follows: \small\begin{verbatim} \Begin(program) \section(sectionname) (content...) . . . \End(program) \end{verbatim}\normalsize \bigskip The main section is \small\verb|\section(regular)|. It is performed to create FORM input for each diagram. The section \small\begin{verbatim} \section(epilog) \end{verbatim}\normalsize will be performed during ``extra call'' after all diagrams have been processed. In the above example this section is used to sum up the final results. The whole TM-program is put into the environment \small\begin{verbatim} \Begin(program,routines.rtn) \End(program) \end{verbatim}\normalsize It contains the argument ``routines.rtn''. This is the name of a file containing various TM-functions and macros, which are loaded in this manner, e.g.\\ \small\verb|\functions()|\normalsize{} -- outputs a comma - separated list of all non-commuting functions occuring in the current diagram and\\ \small\verb|\commuting()|\normalsize{} -- outputs the list of all commuting functions. In our example the global variable ``callform'' is defined by means of the function \small\verb|\ask()|\normalsize{}. This function outputs the argument to the screen waits for the user to reply ``y'' or ``n'' and returns ``Y'' or ``N'', respectively, i.e. ``callform'' then takes the value ``Y'' or ``N''. This action is placed in the environment \small\begin{verbatim} \Begin(initialization) . . . \End(initialization) \end{verbatim}\normalsize and thus is performed only once. The section ``\small\verb|regular|\normalsize'' contains the skeleton of the FORM program. Output, i.e. the FORM input, inside the environment \small\begin{verbatim} \Begin(output,filename) . . . \End(output) \end{verbatim}\normalsize is directed into the file ``filename''. In the above example, instead of an explicit file name, we use the macro \small\verb|\askfilename()|\normalsize{} defined in the style file \small\verb|specmode.tml|\normalsize. This macro asks the filename from the user. For each diagram a filename of the type``\small\verb|d.frm|'' will be expanded to ``\small\verb|d#.frm|'', where \small\verb|#|\normalsize{} is the number of the current diagram. Thus, for diagram number 15 the file ``\small\verb|d15.frm|'' will be created. Note that this expansion is performed by the environment and not by the macro \small\verb|\askfilename()|. The macro \small\verb|\askfilename()|\normalsize{} asks the file name only once in the case when the file name is not defined yet. It is impossible to change the file name once entered. The macro \small\verb|\counter()|\normalsize{} counts the processed diagrams. It can differ from the diagram number (produced by QGRAF for the set of all diagrams) returned by the operator \small\verb|\currentdiagramnumber()|. The operator \small\verb|\blankline()|\normalsize{} returns a blank line. It must be used to obtain a blank line in the produced output because all blank lines in the TM-program are suppressed by default. There is the environment \small\verb|foreach|\normalsize{} providing a ``cycle with parameter''. The content of the environment \small\begin{verbatim} \Begin(foreach,i,1,10) \End(foreach) \end{verbatim}\normalsize will be repeated 10 times. Each time the macro \small\verb|\i()|\normalsize{} expands to the digit counter, i.e 1,2,...10. The built-in operator \small\verb|\numberofinternallines()|\normalsize returns for each diagram the number of internal lines; the built-in operator \small\verb|\maxfcount()|\normalsize{} returns the number of connected fermion lines in the current diagram. The macro \small\verb|\integrand()|\normalsize{} is expanded in a sequence of TM-operators providing the Feynman integrand for the current diagram. There is the environment \small\verb|special|\normalsize. It has the form \small\begin{verbatim} \Begin(special,<condition>) . . . \End(special) \end{verbatim}\normalsize The content of this environment will be performed only if \small\verb|<condition>|\normalsize{} evaluates to \small\verb|True|. In the folowing example this environment is used to call FORM: \small\begin{verbatim} \Begin(special,"\import(callform)"eq"Y") \execute(\(form -l )!.!) \End(special) \end{verbatim}\normalsize The macro \small\verb|\execute()|\normalsize{} expands the symbols \small\verb|!.!|\normalsize{} to the full file name (e.g., \small\verb|d15.frm|\normalsize). The expression \small\verb|!.|\normalsize{} is expanded to the file name without extension, and \small\verb|.!|\normalsize{} is expanded to the extension. The macro \small\verb|\execute()|\normalsize{} calls the built-in TM-operator \small\verb|\system()|\normalsize{} to execute FORM. Thus, for the diagram number 15, e.g., the following command will be fulfilled: \small\begin{verbatim} form -l d15.frm \end{verbatim}\normalsize Here we assume that the user has entered the filename ``d.frm''. The section ``\small\verb|epilog|'' uses the environment ``\small\verb|output|'' without a second argument. The second argument is unnecessary because the file name is already defined. In the ``extracall'' the environment ``\small\verb|output|'' uses the filename exactly the user has specified it, answering the question from the macro \small\verb|\askfilename()|, i.e. ``d.frm'' in this case. \section{Generating the FORM input and executing FORM} DIANA has many command line options. In particular, the user can specify the numbers of the diagrams to be evaluated. The command \small\begin{verbatim} diana -c config.zbb -b 451 -e 453 \end{verbatim}\normalsize tells DIANA to process diagrams number 451, 452 and 453. After executing this command four new files appear in the current directory. If the user has entered the file name ``d.frm'' these files are: ``d451.frm'', ``d452.frm'', ``d453.frm'' and ``d.frm''. The content of the file ``d.frm'' looks like follows: \small\begin{verbatim} #do j = 1,3 load dia'j'.sto; .store #enddo g Rq = #do j = 1,3 + dia'j' #enddo ; .sort print; .end \end{verbatim}\normalsize It can be used by FORM to sum up the calculated diagrams. The file ``d452.frm'' looks like follows: \small\begin{verbatim} **** d2 **** (diagram 452) * Set here your defines! #define mm1 "mmt" #define mm2 "mmt" #define mm3 "0" #define mm4 "0" #define mm5 "0" #define mm6 "mmt" #define LINE "6" #define FERMIONLINE "2" #define TOPOLOGY "top8_" functions F,FF; commuting VV; l Rq = (-1)*F(2,1,mu1,1,0,1)*(-i_)*em*Qd*FF(3,1,-p3,0)*i_* F(3,1,mu3,1,0,1)*(-i_)*em*Qd*F(1,2,mu,GVu,-GAu,1)*(+i_)*em/2/s/c* FF(1,2,-p1,mt)*i_*F(4,2,mu2,1,0,1)*(-i_)*em*Qu*FF(6,2,-p6,mt)*i_* F(5,2,mu4,1,0,1)*(-i_)*em*Qu*FF(2,2,+p2,mt)*i_* VV(4,mu1,mu2,+p4,0)*i_*VV(5,mu3,mu4,+p5,0)*i_; * Here should be your FORM program! .end \end{verbatim}\normalsize From the beginning of this file we can see that the macro \small\verb|\counter()|\normalsize{} produces the order number of the processed diagram, i.e. 2, while the built-in TM operator \small\verb|\currentdiagramnumber()|\normalsize{} produces the number of the current diagram according to the QGRAF file ``qlist.zbb'' i.e. 452. \begin{figure}[ht] \small\begin{verbatim} . . . *--#[ d452: * -1 *vx(B(-2),b(1),A(2)) *vx(B(1),b(-4),A(3)) *vx(T(5),t(4),Z(-1)) *vx(T(4),t(6),A(2)) *vx(T(6),t(5),A(3)) * *--#] d452: . . . \end{verbatim}\normalsize \vskip -47mm \hspace*{7cm}{\epsfysize=40mm \epsfbox{d452.eps}} \vskip 5mm \caption{\label{d452}To the left there is the QGRAF output for the diagram ``number 452''. {\tt t} and {\tt T} stand for t and anit-t quarks, {\tt b} and {\tt B} stand for b and anit-b quarks, {\tt A} is the photon, {\tt Z} is the Z boson. External legs are numerated by negative numbers and {\tt vx} stands for the vertex. The corresponding graph is shown on the right. The lines are numbered corresponding to QGRAF output. The numbers in brackets correspond to the numbering defined by ``topology 8'' in the file ``settings.zbb''. Bold italic digits numerate vertices according to the topology 8.} \end{figure} In Fig. \ref{d452} the QGRAF output for the diagram number 452 is shown together with the corresponding graph. We can see that in this case there are two fermion lines, one corresponding to the external $b-\bar b$, while the second appears as closed fermion loop, the $t$ quark triangle. That is why the \small\verb|#define|\normalsize{} ``FERMIONLINE'' is equal to ``2''. In order to understand the Feynman integrand structure in the above example we consider the simplified SM in use, i.e. the ``gwsmassless.model''. In our example is contained in the file ``settings.zbb''. The following notation is used:\\ {\tt em} -- electromagnetic coupling ($em > 0$);\\ {\tt s} -- $\sin(\theta_W)$;\\ {\tt c} -- $\cos(\theta_W)$;\\ {\tt mt, mW, mZ, mH} -- masses;\\ {\tt mmt, mmW, mmZ, mmH} -- masses squared;\\ {\tt Qe, Qu, Qd} -- electric charges in units of the proton charge, i.e. -1, 1/3, -2/3;\\ {\tt GAe, GAnu, GAu, GAd} -- axial couplings;\\ {\tt GVe, GVnu, GVu, GVd} -- vector couplings; We have two (c)functions for propagators: {\tt FF} and {\tt VV}.\\ Vector propagator:\\ {\tt VV(} \begin{quote} 1.~ number of line\\ 2.~index of the first particle\\ 3.~index of the second particle\\ 4.~type (0 -- photon, 1 -- {\tt Z}, 2 -- {\tt W}, 3--gluon) \end{quote} {\tt )}\\ Fermion propagator:\\ {\tt FF(} \begin{quote} 1.~number of line\\ 2.~fermion line number\\ 3.~ mass \end{quote} {\tt )}\\ For vertices we have one commuting function {\tt V} and one function {\tt F}. \begin{itemize} \item The first argument always is the number of the vertex. \item The last argument shows the type of the vertex: 0 -- scalar, 1 -- vector, 2 -- tensor, 3 -- VVV, 4 -- VVVV \item The function {\tt F} carries as second argument the fermion line number. \end{itemize} For example, the fermion-vector-fermion vertex is:\\ {\tt F(} \begin{quote} 1.~number of line\\ 2.~fermion line number \\ 3.~index\\ 4.~{\tt GV} (vector coupling)\\ 5.~{\tt GA} (axial coupling)\\ 6.~1 (the type of vertex \end{quote} {\tt )} Note that the ``gwsmassless.model'' assumes a zero {\tt b} quark mass (see \small\verb|#define mm3 "0"|\normalsize). \section{Conclusion} Higher order loop calculations and/or multileg ones in one-loop order in the SM of electroweak interactions require the calculation of many complicated Feynman diagrams, which is due to the large particle spectrum of the model. The number of diagrams can exeed many thousands. Therefore the evaluation of such large number of diagrams needs to be automated in almost all respects. First of all an automation of the diagram generation is needed. Since FORM, in our opinion, is the most suitable language for such kind of calculations, we found it of urgent need to write a program for this purpose which directly produces ``FORM input'', i.e. produces an algebraic representation of Feynman diagrams in terms of their momentum representation, written in FORM format. This is what the presented program DIANA is able to do at present and we believe it will have wide applications. For an application of DIANA for the process $Z \to b \bar{b}$ see (\cite{Zbb}). In further steps techniques of handling the numerators and evaluation of scalar Feynman diagrams must be included. These methods partially exist already (\cite{TLAMM,LME}). \vskip 10mm \noindent {\large \bf Acknowledgments} \vskip 10mm \noindent We are grateful to M.~Kalmykov and O.~Veretin for helpful discussions. M.T was supported by Bundesministerium f\"ur Forschung und Technologie under PH/05-7BI92P 9. and in part by RFBR $\#$98-02-16923. \vskip 10mm \noindent
\section{Introduction} Rabies is one of the oldest recorded infectious diseases. Records from the Middle Ages show that rabies was then widespread in western Europe. Throughout the world, dogs are important vectors, however, there exist many early references to several species of wildlife implicated in the spread of the disease~\cite{baer}, \cite{sike}, \cite{mcdo}. Nowadays, most cases of rabies are found among animals (dogs, cats, livestock, wildlife). According to Macdonald~\cite{mcdo}, the present European epizootic (epidemic among animals) began south of Gdansk (Poland) in 1939. Moving westwards, it was first recorded in France in 1968, the Netherlands in 1974, Spain in 1975 and Italy in 1977. Its main vector and victim is the red fox. Rabies is an acute infectious disease of the central nervous system caused by a virus. The disease is transmitted from rabid to susceptible foxes, usually by biting. About half of the infectious foxes become agressive, and lose their sense of direction and territorial behavior, wandering randomly. Murray and co-workers~\cite{kall}, \cite{msb}, \cite{mur} studied different models of the spread of rabies among foxes formulated in terms of partial differential equations. In these models, the fox population is divided into three species: susceptible, infected but non-infectious, and rabid foxes which are infectious. The principal assumptions are: \begin{enumerate} \item[(i)] Susceptibles evolve in time according to a logistic equation. \item[(i)] Susceptibles become infected at an average rate per capita proportional to the number of rabid foxes present. \item[(iii)] Infected foxes become rabid after an average incubation period. \item[(iv)] Rabid foxes die after an average duration of the disease. \item[(v)] Rabid and infected foxes also die of causes other than rabies. \item[(vi)] The spatial spread of the disease is essentially due to the random motion of the rabid foxes \end{enumerate} Such models have unquestionably contributed to our understanding of the spread of an infectious disease, but they do not take correctly into account the short-range character of the infection process, and neglect spatial correlations. In order to include these features, we have built up and studied two lattice models in which the spread of the epidemic is viewed as the growth of a random cluster on a lattice. We will first describe a coupled-map lattice model, and then a two-dimensional automata network model. \section{Lattice models} The spread of an epidemic in a population is a complex process. When modeling such a process, among the many features which are likely to be important, we should, however, only retain the few relevant ones which are thought to play an essential role in the interpretation of the observed phenomena. In both models, at each time step, the fox population evolves according to the following rules: \begin{enumerate} \item[(1)] A susceptible has a probability $b$ to give birth to a susceptible at a nearby empty location. Infected and rabid foxes do not give birth. \item[(2)] Susceptibles and infected have a probability $d$ to die due to natural causes, and $d_\ell$ multiplied by the total local population density of neighboring foxes (\textit{i.e.}\ susceptibles + infected + infectious) due to lack of food. \item[(3)] Susceptibles become infected at a rate proportional to the local density of neighboring rabid foxes, the proportionality factor is the probability $p_i$ to be infected. \item[(4)] Infected, \textit{i.e.} incubating, foxes become rabid with a probability $p_r$. \item[(5)] Non-rabid foxes being territorial, susceptibles and infected foxes evolve without moving to a neighboring territory. \item[(6)] Rabid foxes have a probability $d_r$ to die due to the disease, and $d_\ell$ multiplied by the total local density of neighboring foxes due to lack of food. \item[(7)] Rabid foxes move at random to a neighboring location. \end{enumerate} In order to exhibit some realistic features of the spread of a rabies epidemic, following the analysis of Murray \textit{et al\/}~\cite{msb}, our models have to contain a minimum number of parameters. There is one source term coming from the birth of susceptibles (parameter $b$). To account for the three death processes, natural causes, lack of food, and rabies, we need 3 parameters ($d$, $d_\ell$, and $d_r$). We need to introduce a parameter which measures the probability to be infected by contact (parameter $p_i$). The existence of an incubation period is an essential feature to exhibit decreasing periodic fluctuations following the main wave front of the susceptible population (parameter $p_r$). \subsection{Coupled-map lattice model} A coupled-map lattice is a dynamical system in which space and time are discrete variables while states are continuous~\cite{kane}, \cite{wk}. Here the state of the system is represented by the function $\mathbf{P}:(i,t)\mapsto \mathbf{P}(i,t)$, where $(i,t)\in\mathbb{Z}\times\mathbb{N}$, and $\mathbf{P}\in[0,1]^3$ is a three-dimensional vector whose components are $S(i,t)$, $I(i,t)$, and $R(i,t)$ which denote, respectively, the densities at site $i$ and time $t$ of susceptible, infected and rabid foxes. $\mathbb{Z}$ is the set of all integers and $\mathbb{N}$ is the set of nonnegative integers. In this model, each site corresponds to a specific territory. Susceptibles and infected foxes evolve without moving to a neighboring site. On the contrary, rabid foxes, which have lost their sense of direction, move to one of their two neighboring sites with equal probabilities . According to our assumptions (Rules 1 to 7), the dynamics of the system is governed by the following recurrence relations: \begin{align*} S(i,t+1) &= S(i,t) - d S(i,t) - d_\ell N(i,t)S(i,t)\\ &\qquad + b (1-N(i,t))S(i,t)- p_i R(i,t)S(i,t)\\ I(i,t+1) &= I(i,t) - d I(i,t) - d_\ell N(i,t)I(i,t)\\ &\qquad + p_i R(i,t)S(i,t) - p_r I(i,t)\\ R(i,t+1) &= R(i,t) - d_r R(i,t) - d_\ell N(i,t)R(i,t) + p_r I(i,t)\\ &\qquad + D(R(i-1,t)) + R(i+1,t) - 2 R(i,t)), \end{align*} where $N(i,t)=S(i,t)+I(i,t)+R(i,t)$ is the total fox density at site $i$ and time $t$, and $D$ is the diffusion coefficient characterizing the random motion of the rabid foxes. \subsection{Automata network model} An automata network is a fully discrete dynamical system. That is, space, time and states are discrete variables. In our model, the space consists of a square $L\times L$. Since we are interested in the spread of an epidemic starting a the center of the lattice, we did not choose cyclic boundary conditions. As for a coupled-map lattice, the time variable is a nonnegative integer. The state of the system at time $t$ is described by the function $s_t:(i,j)\mapsto s(i,j;t)$, where $(i,j)\in{\mathbb{Z}}_L^2$, $t\in\mathbb{N}$, and $s(i,j;t)$ can take four values corresponding to the four possible states of the site $(i,j)$ since this site can be either empty or occupied by one of the three fox species (susceptible, infected, rabid). At each time step, the state of the system evolves according to the successive application of the two following subrules: \begin{itemize} \item first a local rule describing birth, death and infection processes, which is applied to all sites synchronously; \item then, a motion rule mimicking rabid fox erratic motion. \end{itemize} Subrule (1) is a probabilistic, two-dimensional, four-state cellular automaton rule, which is defined as follows. At each time step, \begin{enumerate} \item[(i)] each susceptible has a probability $b$ to give birth to a susceptible at an empty first-neighbor site, therefore, the probability that an empty site, having $z_s$ susceptibles among its four first-neighbors, becomes occupied by a susceptible is $1-(1-b)^{z_s}$; \item[(ii)] each susceptible has a probability $d$ to die of natural causes; \item[(iii)] each susceptible having $z_i$ rabid foxes among its four first-neighbors becomes infected with a probability $1-(1- p_i)^{z_i}$; \item[(iv)] each infected has a probability $d$ of die of natural causes; \item[(v)] each infected has a probability $p_r$ to become rabid; \item[(vi)] a rabid fox has a probability $d_r$ to die of the disease; \item[(vii)] each fox (either susceptible, or infected, or rabid) having $z_f$ foxes among its first-neighbors has a probability $d_\ell z_f$ to die due to lack of food. \end{enumerate} Subrule (2) can be described as follows: At time $t$, a rabid fox is selected at random to perform a move to a first-neighbor site also selected at random. If the site is empty the fox will effectively move otherwise it will not. This process is repeated $m N_R(t) L^2$, where $N_R(t)$ denotes the density of rabid foxes at time $t$. The parameter $m$ represents the average number of tentative move per rabid fox at time $t$. Note that this sequential diffusive process allows some rabid foxes to move more than others. Automata networks of this type are called \textit{diffusive cellular automata}. Their general properties have been studied by Boccara \textit{et al\/}~\cite{boc1}. They have been used to build up various epidemic models~\cite{boc2} in which the motion of the individuals play an important r\^ole in the spread of the epidemic. \section{Numerical simulations: results and analysis} \subsection{Coupled-map lattice model} In our simulations, the lattice size is 441. The sites are labeled from $-220$ to $220$. The initial susceptible fox densities have the same value at all sites $i$, while the initial infected and rabid fox densities are nonzero only at the central site. These values are: $S(i,0)=0.6$. $I(0,0)=0.005$, and $R(0,0)=0.005$. For all $t\in\mathbb{N}$, at the boundaries of the chain, the densities of the various fox species satisfy the condition \begin{align*} S(L+1,t) + &I(L+1,t) + R(L+1,t) = \\ &S(-L-1,t) + I(-L-1,t) + R(-L-1,t) = 0 \end{align*} As reported by Macdonald~\cite{mcdo} and Murray~\cite{mur}, the epidemic spread of rabies among foxes is characterized by a traveling epizootic wave front followed by periodic decreasing fluctuations of susceptibles density which tends to its steady state. Our numerical simulations show that our coupled-map lattice model clearly exhibits these features, as illustrated in Figure~1 which represents, at a given time, the variations of susceptible, and rabid fox densities, as a fonction of site location for two different values of $d_\ell$. Increasing $d_\ell$ decreases the height of the peak of the first and subsequent outbreaks of infected and rabid foxes. A smaller value of $d_\ell$ is equivalent to a larger carrying capacity of the environment, therefore, a larger carrying capacity implies a more severe epidemic. Figure~2 illustrates the influence of the diffusion coefficient $D$. As expected, increasing $D$ increases the speed at which the epidemic spreads (see also Figure~3). While the values of the amplitudes of the various fox densities do not change, the distance between two successive outbreaks increases with $D$. According to epidemiological evidence (cf. references in Murray~\cite{mur}) there exists a threshold value for the carrying capacity below which rabies die out. In the case of our model we found that for $d_\ell>0.0456$ the epizootic does not spread (see Figure~4). This threshold depends, of course, upon the values of the other parameters. We have determined numerically the speed of the epizootic wavefront as a function of the diffusion coefficient $D$ for two different values of the parameter $d_\ell$. Our results are represented in Figure~3. As expected from a dimensional argument, this speed varies as $\sqrt{D}$. Increasing $d_\ell$, we verify again that the speed of the epizootic wavefront decreases. We have also determined numerically the speed of the epizootic wavefront as a function of the parameter $d_\ell$. Figure~4 shows that, as we already mentioned, this speed is a decreasing function of $d_\ell$, which goes to zero at a threshold value. \subsection{Automata network model} In our simulations, the lattice size is $201\times 201$. Simulations start from a random initial configuration. The initial densities of susceptible, infected and rabid foxes are, respectively, equal to $0.6$, 0.005, and 0.005. Rabid and infected foxes exist only inside a disk of radius 10 in the initial configuration. Our results are averages over 50 to 100 different initial configurations. As for the coupled-map lattice model, we have studied the influence of the carrying capacity and the diffusion on the various fox species. We found similar results as illustrated in Figures~5a-5c. Increasing $d_\ell$ decreases the height of the peak of the first and subsequent outbreaks of rabid foxes (Figures~5a and 5b), while increasing $m$ increases the speed at which the epidemic spreads (Figures~5b and 5c). Above a threshold value of $d_\ell$, which is equal to 0.053 for $m=0.5$ and 0.0245 for $m=0.1$, the epizootic does not spread (see Figure~9). As time increases, rabid foxes can be found at larger distances from the center of the lattice. For different values of the parameter $m$, we have determined the fractal dimension of the rabid cluster---that is the cluster containing all rabid foxes---as a function of time. If $r_{\max}$ is the radius of this cluster, its fractal dimension $\delta$ is defined by $N_R(t)=\pi r_{\max}^\delta$, where $N_R(t)$ is the number of rabid foxes in the cluster. Figure~6 shows that $\delta$ tends to a constant value as time increases. This limit value increases when $m$ decreases since, as expected, the density of rabid foxes if higher for lower values of $m$. For $d_\ell$ respectively equal to 0.005, 0.01, and 0.015, the corresponding limit values are 1.27, 1.25, and 1.22. Figures~7a-7f show how the cluster containing the infected and infectious foxes grows as a function of time. Infected and infectious foxes are essentially localized at the boundary of the cluster. This feature corresponds to the small subsequent outbreaks following the first one. This is also clearly illustrated in Figures~5. As for the coupled-map lattice model, we have determined numerically how the speed of the epizootic wavefront varies with the carrying capacity and the diffusion. Our results are represented in Figures~8 and 9. The curves representing the variation of the speed of the epizootic wavefront as a function of $d_\ell$ show the existence of a threshold value, which, in particular, depends upon the parameter $m$. \section{Conclusion} We have investigated two different models of the spatial spread of rabies among foxes: a one-dimensional coupled-map lattice model, and a two-dimensional automata network model. In both models, the fox population is divided into three-species: susceptible, infected or incubating, and infectious or rabid. They are based on the fact that susceptible and incubating foxes are territorial while rabid foxes have lost their sense of direction and move erratically out of their territory propagating the disease. coupled-map lattice models and automata networks models have the advantage, compared to models formulated in terms of differential equations, to take into account the short-range character of the infection process. We have essentially studied how the spatial distribution of rabies, and the speed of propagation of the epizootic front depend upon the carrying capacity or food availability of the environment and parameters characterizing the erratic motion of rabid foxes out of their territory. In agreement with ecological studies, our numerical simulations show that, decreasing food availability slows down the spread of the disease, and, that below a certain threshold, rabies eventually dies out. On the other hand increasing the parameters measuring the diffusive motion of rabid foxes favors the spread of the disease. \newpage
\section{Introduction} The recent data of Super-Kamiokande (SK) \cite{sk} have given a new impetus to the atmospheric neutrino problem and a possible interpretation in terms of neutrino oscillation. Moreover the high statistics of SK makes it possible to study the zenith-angle dependence of the neutrino flux from which one can conclude that the $\nu_\mu$'s show signs of oscillation but the $\nu_e$ events are consistent with the no-oscillation hypothesis. Independently the results from the reactor experiment CHOOZ disfavours the $\nu_\mu - \nu_e$ oscillation hypothesis \cite{chooz}. On the other hand large angle $\nu_\mu-\nu_\tau$ or $\nu_\mu-\nu_s$ ($\nu_s$ being a sterile neutrino) solution continues to give a good fit to the data. Nevertheless effort has been on to try out other possibilities to explain the anomaly observed in SK and one among these is neutrino decay \cite{lipari,pak}. In \cite{lipari} it was shown that neutrino decay gives a poor fit to the data. However they considered neutrinos with zero mixing. Barger {\it et al.} considered the situation of neutrino decay in the general case of neutrinos with non-zero mixing angle \cite{pak}. They showed that the neutrino decay fits the $L/E$ distribution of the SK data well. The $\Delta m^2$ taken by them was $>$ 0.1 $eV^2$ so that the $\Delta m^2$ dependent term averages out. As pointed out in \cite{pak} such a constraint on \mbox{$\Delta{m}^{2}$~} is valid when the unstable state decays into some other state with which it mixes. If however the unstable state decays into a sterile state with which it does not mix then there is no reason to assume \mbox{$\Delta{m}^{2}$~} $>0.1~eV^2$. In this paper we present our results of two-flavour $\nu_\mu - \nu_\tau$ oscillation and neutrino decay solutions to the atmospheric neutrino problem by doing $\chi^2$-fit to the 848 days of sub-GeV and multi-GeV Super-Kamiokande data \cite{kate}. We have also presented the results of $\chi^2$-fit to the 535 days SK data and have compared it with the results for the new data. For the neutrino decay analysis we take the most general case of neutrinos with non-zero mixing and consider two pictures \begin{itemize} \item \mbox{$\Delta{m}^{2}$~} $>$ 0.1 $eV^2$ (scenario (a)) \item \mbox{$\Delta{m}^{2}$~} unconstrained (scenario (b)) \end{itemize} We also explicitly demonstrate the behavior of the up-down asymmetry parameters \cite{fl,yasuda} in both scenarios. Our analysis shows that scenario (a) is ruled out at 100\%(99.99\%) C.L. by the 848(535) days of SK data. However if we remove the constraint on $\Delta m^2$ and consider the possibility of decay into a sterile state then one can get an acceptable fit for $\Delta m^2$ $\sim 0.001 eV^2$ and \mbox{$\sin^{2}2\theta$~} large. The plan of the paper is as follows. In section 2 we present our results for two-generation $\nu_\mu-\nu_\tau$ oscillation analysis. In section 3.1 we present our results for the neutrino decay solution constraining \mbox{$\Delta{m}^{2}$~} to be $> 0.1 eV^2$. In section 3.2 we do a three parameter $\chi^2$ analysis by removing the constraint on $\Delta m^2$. In section 4 we perform a comparative study of the three cases and indicate how one can distinguish experimentally between the scenario (b) and the $\nu_\mu - \nu_\tau$ oscillation case though both give almost identical zenith-angle distribution. \section{$\nu_\mu - \nu_\tau$ oscillation} In the two-flavour picture the probability that an initial $\nu_{l}$ of energy $E$ remains a $\nu_{\l}$ after traveling a distance $L$ in vacuum is \begin{equation} P_{\nu_{l}\nu_{l}} =1- \sin{^2}2\theta \sin{^{2}}(\pi L/\lambda_{osc}) \label{p2nu} \end{equation} where $\theta$ is the mixing angle between the two neutrino states in vacuum and $\lambda_{osc}$ is the oscillation wavelength defined as, \begin{equation} \lambda_{osc} = (2.5~ km)~ \frac{E}{GeV} \frac{eV^2}{\Delta m^2} \label{lo} \end{equation} where $\Delta m^2$ denotes the mass squared difference between the two mass eigenstates. The expected number of $l$ (e or $\mu$) like 1 ring events recorded in the detector in presence of oscillations is given by \begin{eqnarray} N_l & = & n_T \int^{\infty}_{0} dE \int^{(E_l)_{\rm max}}_{(E_l)_{\rm min}} dE_l \int_{-1}^{+1} d\cos \psi \int_{-1}^{+1} d\cos \xi\ {1 \over 2\pi} \int_{0}^{2\pi} d\phi \nonumber\\ &\times& {d^2F_l (E,\xi) \over dE~d\cos\xi} \cdot{ d^2\sigma_l (E,E_l,\cos\psi) \over dE_l~d\cos\psi } \epsilon(E_l) \cdot {\ }P_{\nu_l \nu_l} (E, \xi). \label{rate} \end{eqnarray} $n_T$ denotes the number of target nucleons, $E$ is the neutrino energy, $E_l$ is the energy of the final charged lepton, $\psi$ is the angle between the incoming neutrino $\nu_l$ and the scattered lepton $l$, $\xi$ is the zenith angle of the neutrino and $\phi$ is the azimuthal angle corresponding to the incident neutrino direction (the azimuthal angle relative to the $\psi$ has been integrated out). The zenith angle of the charged lepton is then given by \begin{equation} \cos \Theta = \cos \xi \cos \psi + \sin \xi \cos \phi \sin \psi \label{zenith} \end{equation} $d^2F_l /dE d\cos\xi$ is the differential flux of atmospheric neutrinos of type $\nu_l$, $d^2\sigma_l/dE_l d\cos\psi$ is the differential cross section for $\nu_l N \rightarrow l X$ scattering and $\epsilon(E_l)$ is the detection efficiency for the 1 ring events in the detector. The efficiencies that were available to us are not the detection efficiencies of the charged leptons but some function which we call $\epsilon(E)$ defined as \cite{private} \begin{equation} \epsilon(E) = \frac{ \int{ \frac{d\sigma}{dE_l} \epsilon(E_l) dE_l}} {\int{ \frac{d\sigma}{dE_l} dE_l}} \end{equation} $P_{\nu_l \nu_l}$ is the survival probability of a neutrino flavour $l$ after traveling a distance $L$ given by, \begin{equation} L= \sqrt{(R_e+h)^2 - {R_e}^2 \sin^2 \xi} - R_e \cos \xi \end{equation} $R_e$ being the radius of the earth and h is the height of the atmosphere where the neutrinos are produced. We use the atmospheric neutrino fluxes from \cite{honda}. For the sub-GeV events the dominant process is the charged current quasi-elastic scattering from free or bound nucleons. We use the cross-sections given in \cite{gaisser}. The events in multi-GeV range have contributions coming from quasi-elastic scattering, single pion production and multi pion production and we have used the cross-sections given in \cite{lipary}. For the multi-GeV events we assume that the lepton direction $\Theta$ is the same as the incoming neutrino direction $\xi$. But actually they are slightly different. We simulate this difference in the zenith angles by smearing the angular distribution of the number of events with a Gaussian distribution having a one sigma width of $15^{\rm o}$ for $\mu$ type events and $25^{\rm o}$ for the e type events \cite{G_G}. For the sub-GeV events, difference in direction between the charged lepton and the neutrinos are exactly taken care of according to eq. (\ref{rate}) and (\ref{zenith}). To reduce the uncertainty in the absolute flux values the atmospheric neutrino measurements are usually presented in terms of the double ratio \begin{equation} R = {\frac{(\nu_\mu + \overline{\nu}_{\mu})/ (\nu_e + \overline{\nu}_e)_{\rm obsvd}} {(\nu_\mu + \overline{\nu}_{\mu})/(\nu_e + \overline{\nu}_e)_{\rm MC}}} \label{ratm} \end{equation} where {MC} denotes the Monte-Carlo simulated ratio. Different calculations agree to within better than 5\% on the magnitude of this quantity. We use a similar quantity R, where \begin{equation} R \equiv \frac{(N_{\mu}/N_e)|_{osc}}{(N_{\mu}/N_e)|_{no-osc}}. \end{equation} The quantities $N_{e,\mu}$ are the numbers of $e$-like and $\mu$-like events, as per eq.(\ref{rate}). The numerator denotes numbers obtained from eq.(\ref{rate}), while the denominator the numbers expected with the survival probability as 1. At the detector, the neutrino flux come from all directions. Thus, the total path length between the production point in the atmosphere and the detector varies from about 10 km to 13,000 km depending on the zenith angle. Neutrinos with zenith angle less than $90^{\rm o}$ ({\it `downward neutrinos'}) travel a distance of $\sim$ 10 -- 100 km from their production point in the atmosphere to the detector while the neutrinos with larger zenith angles ({\it `upward neutrinos'}) cross a distance of up to $\sim$ 13,000 km to reach the detector. Apart from altering the flavour-content of the atmospheric neutrino flux, oscillations could lead to the following effect: if the oscillation length is much longer than the height of the atmosphere but smaller than the diameter of the earth, only upward neutrinos coming from the opposite side of the earth will have significant oscillations. These would show up as an up-down asymmetry in the event distribution. SK has enough statistics to study these up-down flux asymmetry. They divide the $(-1,+1)$ interval in $\cos\Theta$ in five equal bins: $(-1.0,-0.6)$, $(-0.6,-0.2)$, $(-0.2,+0.2)$, $(+0.2,+0.6)$, $(+0.6,+1.0)$ and give the number of events in each bin. The first two bins correspond to the upward neutrinos and the last two bins correspond to the downward neutrinos. To probe the up-down flux asymmetries we use the parameter $Y$ defined in \cite{yasuda}, \begin{equation} Y_{l} \equiv {(N_{l}^{-0.2}/N_{l}^{+0.2})|_{osc} \over (N_{l}^{-0.2}/N_{l}^{+0.2})|_{no-osc}}. \end{equation} Here $N_{l}^{-0.2}$ denotes the number of $l$-type events produced in the detector with zenith angle $\cos \Theta < -0.2$, {\it i.e.} the upward neutrino events while $N_{l}^{+0.2}$ denotes the number of $l$-type events for $\cos \Theta > 0.2$ {\it i.e.} events coming from downward neutrinos. The central bin has contributions from both upward and downward neutrinos and is not useful for studying the up-down asymmetry. We minimize the $\chi^2$ function defined as \cite{yasuda} \begin{equation} \chi^2 = \sum_{i} \left[\left({R}^{exp} - R^{th} \over \delta R^{exp} \right)^2 + \left({Y^{exp}_{\mu} - Y^{th}_{\mu} \over \delta Y^{exp}_{\mu}}\right)^2 + \left({Y^{exp}_{e} - Y^{th}_{e} \over \delta Y^{exp}_{e}}\right)^2 \right], \label{chi} \end{equation} where the sum is over the sub-GeV and multi-GeV cases. The experimentally observed rates are denoted by the superscript "exp" and the theoretical predictions for the quantities are labeled by "th". $\Delta R^{exp}$ is the error in $R$ obtained by combining the statistical and systematic errors in quadrature. $\Delta Y^{exp}$ corresponds to the error in $Y$. For this we take only the statistical errors since these are much larger compared to the systematic errors. We include both the $e$-like and the $\mu$-like up-down asymmetries in the fit so that we have 4 degrees of freedom (6 experimental data - 2 parameters) for the oscillation analysis in the two parameters $\Delta m^2$ and $\sin^2 2\theta$. The use of these type of ratios for the $\chi^2$ analysis test has been questioned in \cite{lisi1} because the error distribution of these ratios is non-Gaussian in nature. The alternative is to use the absolute number of e or $\mu$ type events taking into account the errors and their correlations properly \cite{G_G,lisi2}. However as has been shown in \cite{yasuda} the use of the $R$'s and $Y$'s as defined above is justified within the 3$\sigma$ region around the best-fit point for a high statistics experiment like SK and provides an alternative way of doing the $\chi^2$-analysis. A comparison of the results of \cite{yasuda} with those obtained in \cite{G_G,lisi2} shows that the best-fit points and the allowed regions obtained do not differ significantly in the two approaches of data fitting. The advantage of using the ratios is that they are relatively insensitive to the uncertainties in the neutrino fluxes and cross-sections as the overall normalization factor gets canceled out in the ratio. We have included the $Y_e$ in our analysis because to justify the $\nu_\mu - \nu_\tau$ oscillation scenario, it is necessary to check that $\chi^2$ including the data on electron events gives a low value and hence it is the standard practice to include these in the $\chi^2$-analysis \cite{yasuda,G_G,lisi2}. The data that we have used are shown in Table 1 which corresponds to the 848 days \cite{kate} and the 535 days \cite{data} of data. \begin{description} \item{Table 1:} The SK data used in this analysis. \end{description} \[ \begin{array}{|c|c|c|c|c|} \hline {} & \multicolumn {2}{c|} {\rm 848 ~ days ~ data} & \multicolumn {2}{c|} {\rm 535 ~ days ~ data} \\ \cline {2-5} {\rm Quantity} & {\rm Sub-GeV} & {\rm Multi-GeV} & {\rm Sub-GeV} & {\rm Multi-GeV} \\ \hline {R^{exp}} & {0.69} & {0.68} & {0.63} & {0.65} \\ \hline {\Delta R^{exp}} & {0.05} & {0.09} & {0.06} & {0.09} \\ \hline {Y^{exp}_\mu} & {0.74} & {0.53} & {0.76} & {0.55} \\ \hline {\Delta{Y^{exp}_\mu}} & {0.04} & {0.05} & {0.05} & {0.06} \\ \hline {Y^{exp}_e} & {1.03} & {0.95} & {1.14} & {0.91} \\ \hline {\Delta {Y^{exp}_e}} & {0.06} & {0.11} & {0.08} & {0.13} \\ \hline \end{array} \] For the 2 flavour ${\nu_{\mu}}-{\nu_{\tau}}$ oscillation the $\chi^2_{min}$ that we get is 1.21 with the best-fit values as \mbox{$\Delta{m}^{2}$~} = 0.003 $eV^2$ and \mbox{$\sin^{2}2\theta$~} = 1.0. This provides a good fit to the data being allowed at 87.64\% C.L. If we use the 535 days data then the $\chi^2_{min}$ that we get is 4.25 with the best-fit values as \mbox{$\Delta{m}^{2}$~} = 0.005 $eV^2$ and \mbox{$\sin^{2}2\theta$~} = 1.0, the g.o.f being 37.32\%. Thus the fit becomes much better with the 848 days data with no significant change in the best-fit values. Though we have used a different procedure of data fitting, our results agree well with that obtained by the SK collaboration\footnote{The best-fit values that the SK collaboration has got for the 848 days data are \cite{kate} \mbox{$\Delta{m}^{2}$~} = 0.003 eV$^2$, \mbox{$\sin^{2}2\theta$~} = 0.995 and $\chi^2/d.o.f$ = 55.4/67. This corresponds to a g.o.f of 84.33\%.}. In fig. 1 we show the 90\% C.L. ($\chi^2 \leq \chi^2_{min} + 4.61$) and the 99\% C.L. ($\chi^2 \leq \chi^2_{min} + 9.21$) allowed region in the (\mbox{$\Delta{m}^{2}$~}, \mbox{$\sin^{2}2\theta$~}) plane for the $\nu_\mu - \nu_\tau$ oscillation hypothesis using the latest SK data. \section{Neutrino decay} The neutrino decay hypothesis assumes that there is an unstable component in ${\nu_{\mu}}$ (say $\nu_2$) which decays into one of the lighter states (say $\nu_3$). Experimental considerations constrain ${\nu_e}$ to decouple from $\nu_2$ and it's decay partners, so that \begin{equation} {\nu_e} \approx \nu_1 \end{equation} \begin{equation} \nu_\mu \approx \nu_2 \cos\theta + \nu_3 \sin\theta \label{nm} \end{equation} From (\ref{nm}) the survival probability of the ${\nu_{\mu}}$ of energy E, with an unstable component $\nu_2$, after traveling a distance $L$ is given by, \begin{eqnarray} P_{{\nu_{\mu}}\numu} &=& \sin^4\theta + \cos^4\theta \exp(-4 \pi L/\lambda_d)\nonumber\\ && {}+ 2\sin^2\theta \cos^2\theta \exp(-2 \pi L/\lambda_d) \cos(2 \pi L/\lambda_{osc})\,, \label{pmumudo} \end{eqnarray} where $\lambda_d$ is the decay length (analogous to the oscillation wavelength given by eq. (\ref{lo})) defined as, \begin{equation} \lambda_d = 2.5 km \frac{E}{GeV} \frac{eV^2}{\alpha} \label{ld} \end{equation} and $\alpha = m_2/\tau_0$, $m_2$ being the mass of the state $\nu_2$ and $\tau_0$ the decay lifetime. The $\lambda_{osc}$ appearing in eq. (\ref{pmumudo}) is the wavelength of oscillations as defined in eq. (\ref{lo}) with \mbox{$\Delta{m}^{2}$~} = $m_2^2 - m_3^2$. \subsection{$\Delta m^2 > 0.1 eV^2$} If the unstable component in the $\nu_\mu$ state decays to some other state with which it mixes then bounds from $K$ decays imply $\Delta m^2 > 0.1 eV^2$ \cite{pak2}. In this case the $\cos(2 \pi L/\lambda_{osc})$ term averages to zero and the probability becomes \begin{equation} P_{{\nu_{\mu}}\numu} = \sin^4\theta + \cos^4\theta \exp(-4 \pi L/\lambda_d) \,. \label{pmumu} \end{equation} In figs. 2 and 3 we show the variation of $R$ and $Y$ with $\alpha$ for various values of $\sin^2 \theta$ for the sub-GeV and multi-GeV cases. For higher values of $\alpha$, the decay length $\lambda_d$ given by eq. (\ref{ld}) is low and the exponential term in the survival probability is less implying that more number of neutrinos decay and hence $R$ is low. As $\alpha$ decreases the decay length increases and the number of decaying neutrinos decreases, increasing $R$. For very low values of $\alpha$ the exponential term goes to 1, the neutrinos do not get the time to decay so that the probability becomes $1-\frac{1}{2}\sin^2 2\theta$ and remains constant thereafter for all lower values of $\alpha$. This is to be contrasted with the $\nu_\mu-\nu_\tau$ oscillation case where in the no oscillation limit the $\sin^2 (\pi L/\lambda_{osc})$ term $\rightarrow$ 0 and the survival probability $\rightarrow$ 1. For multi-GeV neutrinos since the energy is higher the $\lambda_d$ is higher and the no decay limit is reached for a larger value of $\alpha$ as compared to the sub-GeV case. This explains why the multi-GeV curves become flatter at a higher $\alpha$. The behavior of the up-down asymmetry parameter is also completely different from the only oscillation case \cite{yasudafig}. In particular the plateau obtained for a range of \mbox{$\Delta{m}^{2}$~} which was considered as a characteristic prediction for up-down asymmetries is missing here. For the decay case even for $\alpha$ as high as 0.001 $eV^2$, the decay length $\lambda_d = 2500~(E/GeV)~km$ so that the exponential term is 1, there is almost no decay for the downward neutrinos and the survival probability is $P = 1 - \frac{1}{2}\sin^2 2\theta$ while the upward going neutrinos have some decay and so $Y$ is less than 1. As $\alpha$ decreases, the $\lambda_d$ increases, and the fraction of upward going neutrinos decaying decreases and this increases $Y$. For very small values of $\alpha$ even the upward neutrinos do not decay and $Y \rightarrow$ 1 being independent of $\theta$. We also perform a $\chi^2$ analysis of the data calculating the "th" quantities in (\ref{chi}) for this scenario. The best-fit values that we get are $\alpha = 0.33 \times 10^{-4}$ in $eV^2$ and $\sin^2\theta = 0.03$ with a $\chi^2_{\rm{min}}$ of 49.16. For 4 degrees of freedom this solution is ruled out at 100\% C.L. The best-fit values for the 535 days of data that we get are $\alpha = 0.28 \times 10^{-4}$ in $eV^2$ and $\sin^2\theta = 0.08$ with a $\chi^2_{\rm{min}}$ of 31.71. For 4 degrees of freedom this solution is ruled out at 99.99\% C.L. \cite{lisidecay}. Thus the fit becomes worse with the 848 days data as compared to the 535 days data. We have marked the $R$ and $Y$ corresponding to the best-fit value of the parameters $\alpha$ and $\sin^2\theta$ in figs. 2 and 3. It can be seen that the best-fit value of R for the sub-GeV neutrinos is just below and that for the multi-GeV neutrinos is just above the $\pm 1\sigma$ allowed band of the SK 848 days of data. The up-down asymmetry parameter $Y$ is quite low for the sub-GeV neutrinos and extremely high for the multi-GeV neutrinos as compared to that allowed by the data. The fig. 2 shows that for the sub-GeV neutrinos the data demands a lower value of $\alpha$ while from fig. 3 we see that the multi-GeV neutrinos need a much higher $\alpha$ to explain the SK data. It is not possible to get an $\alpha$ that can satisfy both the sub-GeV and the multi-GeV SK data, particularly it's zenith angle distribution. In this scenario, decay for the sub-GeV upward neutrinos is more than that for the multi-GeV upward neutrinos (downward neutrinos do not decay much) and as a result $Y$ for sub-GeV is lower than the $Y$ for multi-GeV, a fact not supported by the data. Since the 848 days data needs even lesser depletion of the sub-GeV flux as compared to the multi-GeV flux, the fit gets worse. \subsection{$\Delta m^2$ unconstrained} In this section we present the results of our $\chi^2$-analysis removing the constraint on $\Delta m^2$. This case corresponds to the unstable neutrino state decaying to some sterile state with which it does not mix \cite{pak}. The probability will be still given by eq. (\ref{pmumudo}). In fig. 4 and 5 we plot the R vs. \mbox{$\Delta{m}^{2}$~} and Y vs. \mbox{$\Delta{m}^{2}$~} for the sub-GeV and multi-GeV data for $\alpha$ = 0.3 $\times 10^{-5} eV^2$ (which is the best-fit value we get for the 848 days data) and compare with the curve obtained for the best-fit value of $\sin^2 \theta$ (=0.5) for the only oscillation case (solid line). For the best-fit value of $\alpha$ that we get, the downward neutrinos do not have time to decay while the upward neutrinos undergo very little decay. Thus the curves are very similar in nature to the only oscillation curves. In the sub-GeV case (fig. 4), for high values of \mbox{$\Delta{m}^{2}$~} around 0.1 $eV^2$ both upward and downward neutrinos undergo \mbox{$\Delta{m}^{2}$~} independent average oscillations and R stays more or less constant with \mbox{$\Delta{m}^{2}$~}. For the upward going neutrinos in addition to average oscillation there is little amount of decay as well and hence Y $\sim$ $N_{up}/N_{down}$ is $\stackrel{<}{\sim}$ 1. As \mbox{$\Delta{m}^{2}$~} decreases to about 0.05 $eV^2$ the oscillation wavelength increases -- for upward neutrinos it is still average oscillation but for the downward neutrinos, the $\cos (2\pi L/\lambda_{osc})$ term becomes negative which corresponds to maximum oscillation effect and the survival probability of these neutrinos decreases, and hence R decreases; while the upward neutrinos continue to decay and oscillate at the same rate and $Y$ becomes greater than 1. As \mbox{$\Delta{m}^{2}$~} decreases further, the downward neutrino oscillation wavelength becomes greater than the distance traversed and they are converted less and less and thus R increases and Y decreases. Below \mbox{$\Delta{m}^{2}$~} = 0.001 $eV^2$ the downward neutrinos stop oscillating completely while for the upward neutrinos the $\cos (2\pi L/ \lambda_{osc})$ term goes to 1, and R and Y no longer vary with \mbox{$\Delta{m}^{2}$~}. For the multi-GeV case (fig. 5) the oscillation wavelength is more than the sub-GeV case and for \mbox{$\Delta{m}^{2}$~} around 0.1 $eV^2$ the $\cos (2 \pi L/\lambda_{osc})$ term stays close to 1 for the downward neutrinos; while the upward neutrinos undergo average oscillations and slight decay and Y is less than 1. As \mbox{$\Delta{m}^{2}$~} decreases the downward neutrinos oscillate even less and the upward neutrinos also start departing from average oscillations and hence $R$ increases and $Y$ decreases. Around 0.01 $eV^2$ the downward neutrinos stop oscillation while for upward neutrinos the oscillation effect is maximum ($\lambda \sim L/2$) and the $cos (2\pi L/ \lambda_{osc})$ term is $\sim$ -1 and $Y$ stays constant with \mbox{$\Delta{m}^{2}$~}. As \mbox{$\Delta{m}^{2}$~} decreases further the upward neutrino oscillation wavelength increases and they oscillate less in number making both R and Y approach 1 for \mbox{$\Delta{m}^{2}$~} around 0.0001 $eV^2$. For multi-GeV neutrinos the decay term contributes even less as compared to the sub-GeV case. We perform a $\chi^2$ minimization in the three parameters \mbox{$\Delta{m}^{2}$~}, \mbox{$\sin^{2}2\theta$~} and $\alpha$. The best-fit values that we get are $\Delta m^2 = 0.003 eV^2$, \mbox{$\sin^{2}2\theta$~} = 1.0 and $\alpha = 0.3 \times 10^{-5} eV^2$. The $\chi^2$ minimum that we get is 1.11 which is an acceptable fit being allowed at 77.46\% C.L.. For the 535 days data the best-fit values that we get are $\Delta m^2 = 0.002 eV^2$, \mbox{$\sin^{2}2\theta$~} = 0.87 and $\alpha = 0.0023 eV^2$ with a $\chi^2_{\rm {min}}$ of 4.14 which is allowed at 24.67\% C.L.. Thus compared to the 535 days data, the fit improves immensely and the best-fit shifts towards the oscillation limit, the best-fit value of the decay constant $\alpha$ being much lower now. It is to be noted however, that the best-fit in this model does not come out to be $\alpha = 0.0$, {\it viz} the only oscillation limit. In table 2 we give the contributions to $\chi^2$ from the $R$'s and $Y$'s at the best-fit value of $\alpha$ and for the $\alpha$ = 0.0 case. \begin{description} \item{Table 2:} The various contributions to the ${\chi^2}_{\rm{min}}$ at the best-fit value of $\alpha$ and at $\alpha$=0.0 \end{description} \[ \begin{array}{|c|c|c|} \hline {\rm Quantity} & {\rm \alpha = 0.3 \times 10^{-5}~eV^2} & {\rm \alpha = 0.0 ~eV^2} \\ \hline {R^{sg}} & {0.085} & {0.021} \\ \hline {Y^{sg}_{\mu}} & {0.011} & {0.033} \\ \hline {R^{mg}} & {0.48} & {0.56} \\ \hline {Y^{mg}_\mu} & {0.014} & {0.073} \\ \hline {Y^{sg}_e} & {0.344} & {0.344} \\ \hline {Y^{mg}_e} & {0.176} & {0.176} \\ \hline \end{array} \] Thus from the contributions to $\chi^2$ we see that for the best- fit case there is improvement for the multi-GeV R and Y as compared to the $\alpha = 0.0$ case. The $\chi^2$ for sub-GeV Y also improves. In fig. 6 we plot the $\chi^2 - \chi^2_{\min}$ vs. $\alpha$ with \mbox{$\Delta{m}^{2}$~} and \mbox{$\sin^{2}2\theta$~} unconstrained. There are two distinct minima in this curve -- one for lower values and another at higher values of $\alpha$. The best-fit \mbox{$\Delta{m}^{2}$~} in both cases is $\sim$ 0.001 $eV^2$. In this model there are two competing processes -- oscillation and decay. For lower values of $\alpha$ the decay length is greater than the the oscillation wavelength and oscillation dominates. The decay term exp$(-\alpha L/E)$ is close to 1 and does not vary much with the zenith distance $L$. As $\alpha$ increases the exponential term starts varying very sharply with $L$ and the variation is much more sharp for the sub-GeV as compared to multi-GeV. This behavior is inconsistent with the data and that is why one gets a peak in $\Delta \chi^2$ for higher $\alpha$. As $\alpha$ increases further the exp$(-\alpha L/E)$ term goes to zero for the upward neutrinos and there is complete decay of these neutrinos while the downward neutrinos do not decay, the exponential term still being 1. Whenever the exponential term is 0 or 1 for the upward neutrinos, the wrong energy dependence of this term does not spoil the fit and these scenarios can give good fit to the data. Even though fig. 6 shows that the data allows a wide range of $\alpha$, we get the two distinct minima in the $\Delta \chi^2$ vs. $\alpha$ curve for high and low $\alpha$ values, for both the 535 days (dotted line) and 848 days (solid line) data. But while the 848 days data prefers the lower $\alpha$ limit, the 535 days data gives a better fit for the high $\alpha$ limit. The reason behind this is that for the 848 days data the R is much higher than for the 535 days data. Hence the 848 days data prefers lower $\alpha$ and hence lower suppression. In fig. 7 we show the 90\% and 99\% C.L. allowed parameter region in the \mbox{$\Delta{m}^{2}$~}- \mbox{$\sin^{2}2\theta$~} plane for a range of values of the parameter $\alpha$. In fig. 8 we show the 90\% and 99\% C.L. contours in the $\alpha$ - \mbox{$\sin^{2}2\theta$~} plane fixing $\Delta m^2$ at different values. These contours are obtained from the definition $\chi^2 \leq {\chi^2}_{\min} + \Delta \chi^2$, with $\Delta \chi^2$ = 6.25 and 15.5 for the three parameter case for 90\% and 99\% C.L. respectively. The bottom left panel in fig 7 is for the best-fit value of $\alpha$. For high $\alpha$ (the top left panel) no lower limit is obtained on $\Delta m^2$, because even if \mbox{$\Delta{m}^{2}$~} becomes so low so that there is no oscillation the complete decay of upward neutrinos can explain their depletion. As we decrease $\alpha$ the allowed parameter region shrinks and finally for $\alpha=0$ we get the two parameter limit modulo the small difference in the C.L. definitions for the two and three parameter cases. The upper right panel of fig. 8 corresponds to the best-fit value of \mbox{$\Delta{m}^{2}$~}. For very low $\alpha$, even though there is no decay, we still have oscillations and that ensures that when \mbox{$\Delta{m}^{2}$~} is large enough there is no lower bound on $\alpha$ as evident in the fig. 8. For \mbox{$\Delta{m}^{2}$~}$=10^{-4} eV^2$ the neutrinos stop oscillating and hence we get a lower bound on $\alpha$ beyond which the depletion in the neutrino flux is not enough to explain the data. \section{Comparison and Conclusion} In fig. 9 we show the histogram of the muon event distributions for the sub-GeV and multi-GeV data under the assumptions of $\nu_\mu-\nu_\tau$ oscillation, and the two scenarios of neutrino decay for the best-fit values of the parameters both for the 535 and the 848 days of data. From the figs. it is clearly seen that the scenario (a) (big dotted line) ($\mbox{$\Delta{m}^{2}$~} > 0.1 eV^2$) does not fit the data well there being too much suppression for the sub-GeV upward going neutrinos and too less suppression for the multi-GeV upward going neutrinos. The scenario (b) (\mbox{$\Delta{m}^{2}$~} unconstrained, small dashed line), however, reproduces the event distributions well. However with the 848 days data the sub-GeV events are reproduced better as compared to the 535 days data and the quality of the fit improves. The neutrino decay is an interesting idea as it can preferentially suppress the upward $\nu_\mu$ flux and can cause some up-down asymmetry in the atmospheric neutrino data. However the intrinsic defect in the decay term $\exp(-\alpha L/E)$ is that one has more decay for lower energy neutrinos than for the higher energy ones. Thus neutrino decay by itself fails to reproduce the observed data \cite{lipari}. If however one considers the most general case of neutrinos with non-zero mixing then there are three factors which control the situation \begin{itemize} \item the decay constant $\alpha$ which determines the decay rate \item the mixing angle $\theta$ which determines the proportion of neutrinos decaying and mixing with the other flavour \item the \mbox{$\Delta{m}^{2}$~} which determines if there are oscillations as well \end{itemize} If the heavier state decays to a state with which it mixes then \mbox{$\Delta{m}^{2}$~} has to be $> 0.1 eV^2$ because of bounds coming from $K$ decays \cite{pak2}. The best-fit value of $\alpha$ that one gets is $0.33\times 10^{-4} eV^2$ with the latest SK data. At this value of $\alpha$ the $e^{-\alpha L/E}$ term tends to 1 for the downward going neutrinos signifying that they do not decay much. The survival probability goes to ($1 - \frac{1}{2} \sin^2 2 \theta $) which is just the average oscillation probability. In order to suppress this average oscillation the best-fit value of $\sin^2 \theta$ comes out to be small in this picture. For the upward going neutrinos, in scenario (a), there will be both decay and average oscillations. If one had only average oscillation then the probability would have stayed constant for a fixed value of the mixing angle $\theta$. But because of the exponential decay term the survival probability drops very sharply as we go towards $\cos\Theta$=-1.0. The drop and hence the decay is more for lower energy neutrinos. As a result the sub-GeV flux gets more depleted than the multi-GeV flux, a fact not supported by the data. In fact the 848 days data requires the sub-GeV flux to be even less suppressed than the multi-GeV flux as compared to the 535 days data and the fit worsens with the 848 days data. The small mixing signifies that the $\nu_\mu$ has a large fraction of the unstable component $\nu_2$ (see eq. (\ref{nm})). Hence the constant $\alpha$ comes out to be low so that the decay rate is less to compensate this. However even at the best-fit $\alpha$ of 0.33 $\times 10^{-4} eV^2$ the survival probability in the bin with $\cos \Theta$ between -1.0 to -0.6 comes out to be 0.15 for E=1 GeV, much lower than the value of $\sim$ 0.5 as required by the data. Thus scenario (a) fails to explain the upward going neutrino data properly because of two main reasons \begin{itemize} \item{$\theta$ is low in order to suppress the average oscillations of the downward neutrinos} \item{the energy dependence of the exponential decay term is in conflict with the data} \end{itemize} In the scenario (b), in addition to mixing with $\nu_\tau$, the unstable component in $\nu_\mu$ decays to some sterile state with which it does not mix. In this case there is no restriction on \mbox{$\Delta{m}^{2}$~} and it enters the $\chi^2$ fit as an independent parameter. We find that \begin{itemize} \item{ The best-fit \mbox{$\Delta{m}^{2}$~} does not come out naturally to be in the \mbox{$\Delta{m}^{2}$~} independent average oscillation regime of $>$ 0.1 $eV^2$, rather it is $0.003 eV^2$.} \item{The best-fit value of the decay constant $\alpha = 0.3 \times 10^{-5} eV^2$ implying that the decay rate is small so that the mixing angle is maximal ($\sin^2 \theta = 0.5$).} \item{Large values of $\alpha$ giving complete decay of upward neutrinos are also allowed with a high C.L. In fact with 535 days data the best-fit was in this region.} \item{ The best-fit value of the decay constant $\alpha$ is non-zero signifying that a little amount of decay combined with \mbox{$\Delta{m}^{2}$~} dependent oscillations gives a better fit to the data.} \end{itemize} At the best-fit values of the parameters there is no oscillation of the downward neutrinos so that the $\cos (2 \pi L/\lambda_{osc})$ term goes to 1. The decay term also goes to 1 signifying that there is not much decay either for the downward neutrinos and the survival probability is $\approx$ 1 without requiring the mixing angle to be low. On the other hand for the upward neutrinos there are oscillations as well as little amount of decay. The sub-GeV upward neutrinos have smaller oscillation wavelength and they are close to the average oscillation limit (survival probability $\sim$ 0.5) while for the multi-GeV neutrinos the oscillation wavelength is such that one has maximum oscillations and the survival probability is less than 0.5. Thus this scenario reproduces the correct energy dependence of the suppression -- namely sub-GeV is suppressed less as compared to multi-GeV neutrinos. The best-fit value of $\alpha$ being even smaller now than the scenario (a) the decay term $e^{-\alpha L/E}$ does not vary very sharply with the zenith distance $L$ or the energy $E$ so that its wrong energy dependence does not spoil the fit. The conversion probability of ${\nu_{\mu}}$ to ${\nu_{\tau}}$ is given by \begin{equation} P_{{\nu_{\mu}}{\nu_{\tau}}} = \frac{1}{4} \sin^2 2\theta\{1+\exp(-4\pi L/\lambda_d) -2 \exp(-2\pi L/\lambda_d) \cos(2\pi L/\lambda_{osc})\} \label{ntb} \end{equation} The value of $P_{{\nu_{\mu}}{\nu_{\tau}}}$ integrated over the energy and the zenith angle, for $\alpha=0.3\times 10^{-5} eV^2$ (the best fit for scenario (b)) is 0.33 for sub-GeV and 0.26 for multi-GeV. For $\alpha=0.44\times 10^{-3} eV^2$ (the second minima in the $\Delta \chi^2$ vs. $\alpha$ curve) the corresponding numbers are 0.21 and 0.15, while for the only ${\nu_{\mu}}-{\nu_{\tau}}$ oscillation case, the corresponding values are 0.37 and 0.26 respectively. The value of \mbox{$\Delta{m}^{2}$~} and \mbox{$\sin^{2}2\theta$~} for both the cases is $0.003 eV^2$ and 1.0 respectively. The fig. 9 shows that the zenith angle dependence of the scenario (b) is almost similar to the case of $\nu_\mu - \nu_\tau$ oscillation. But the two cases are very different in principle. For the oscillation case a larger $\theta$ implies a larger conversion whereas in scenario (b) a larger $\theta$ means the fraction of the unstable component is less in $\nu_\mu$ and the depletion is less. If one compares the conversion probability as given by eq.(\ref{ntb}) with the one for the $\nu_\mu - \nu_\tau$ oscillation case, then the scenario (b) considered in this paper would have smaller number of $\nu_\tau$s in the resultant flux at the detector, especially for the larger values of $\alpha$ which are still allowed by the data and the two cases might be distinguished when one has enough statistics to detect $\tau$ appearance in Super-Kamiokande \cite{hall} or from neutral current events \cite{smirnov}. In our paper we have followed the procedure of data fitting as done in \cite{yasuda}. Thus we use the ratios for which the common systematic errors get canceled out. Strictly speaking one should use the absolute number of events and include all the correlations between bins and $e$-like and $\mu$-like events. But the best-fit points and the allowed regions are not expected to change significantly \cite{private}. We have compared the scenarios of neutrino oscillation and decay with the same definition of $\chi^2$ and for this purpose of comparison neglecting the correlation matrix will not make much difference. Apart from the statistical analysis we have given plots of $R$ and $Y$ for various values of the parameters. The allowed parameter ranges from these plots are consistent with what we get from our statistical analysis. The histograms that we have plotted are also independent of our definition of $\chi^2$. We have checked that if we estimate the allowed ranges from the histograms these are consistent with what we get from our definition of $\chi^2$. Thus we agree with the observation in ref. \cite{yasuda} that although this method of data fitting is approximate it works well. In our analysis we have used only the SK data because it has the highest statistics as compared to the earlier atmospheric neutrino experiments. Radiative decays of neutrinos are severely constrained \cite{fukugita} and what we consider here are the non-radiative decay modes of the neutrino. Models for neutrino decay for the scenario (a) are discussed in \cite{pak,anjan}. For the scenario (b) the unstable state decays to a sterile neutrino state and a light scalar. The model described in \cite{acker} in connection to the solar neutrino problem can be adapted to emulate this scenario. A recent paper \cite{pakvasa2} has discussed how such a model can be constructed. Since in scenario (b) the decay products are invisible there are no distinctive signs of the decay. Decay of leptons to the light scalar are prohibited from conservation of lepton number. Hence it is difficult to constrain these from laboratory experiments \cite{acker}. Consequences of such a model for astrophysics has been discussed in \cite{pakvasa2}. \vspace{1cm} We would like to thank Anjan Joshipura for creating our interest in the neutrino decay solution to the atmospheric neutrino problem. We also like to thank Osamu Yasuda for many useful correspondences during the development of our computer code, S. Pakvasa for useful correspondences and Kamales Kar for discussion and encouragement. Finally we would like to thank Kate Scholberg for providing us with the 848 days data and for other useful correspondences.
\section{Introduction} It is widely believed that the standard model (SM) of electroweak interactions \cite{SM}, although so far in beautiful agreement with experimental data, is not a final theory, but rather an effective low energy theory\footnote{% For a recent review and complete list of references on the application of the effective lagrangian approach within the theory of the electroweak interaction see e.g. \cite{Wudka} and references therein. For the general principles of the effective lagrangians see also \cite{Georgi}.}, valid for energies well below some scale $\Lambda $, which characterizes the onset of a new physics. This scale could be e.g. the mass of yet unobserved heavy particles present in the spectrum of the more fundamental theory. Integrating out these high energy degrees of freedom, one ends up with an effective lagrangian which describes only the interactions of the particles belonging to the low energy part of the original spectrum. Such an effective lagrangian should reflect the most general features of the original theory such as various types of symmetries, modes of their realization, anomaly matching conditions etc. On the other hand it is not constrained to be renormalizable in the usual sense, i.e. it can (and in general it has to, in order to be consistent) contain interaction terms with canonical dimension larger then four. These are multiplied with coupling constants, proportional to the inverse powers of some mass parameter related to the scale $\Lambda $% , and therefore suppressed with respect to the renormalizable terms (the SM should be understood within this framework as the lowest order part of the complete effective lagrangian within the expansion in the powers of $\Lambda ^{-1}$ ).The low energy coupling constants (LEC) could be at least in principle calculable (e.g. by integrating the heavy degrees of freedom order by order in perturbation theory and imposing necessary matching conditions at the threshold of new physics) provided the fundamental theory were known. However in the most of applications of the effective lagrangian approach this is not the case either because the explicit calculations are not possible, or, as in the case of SM, the fundamental theory of the new physics is not known. Rather one lets these parameters a priori undetermined and treat them as a useful parameterization of the dynamics of the yet unknown fundamental theory. Relaxing the constraint of the renormalizability of the interactions does not mean at all the complete loss of predictivity, which was the traditional argument for rejecting such theories. As a rule, within the framework of the effective theories there are well defined expansion prescriptions, which enable one to calculate loops and to absorb the infinities in the finite number of renormalized LEC at each order. These may be in principle measured experimentally and then used as an input for other predictions. Going to higher orders in the expansion, the number of LEC increases considerably, however their importance decreases because of suppression by negative powers of $\Lambda $. One can adopt also another point of view and use the measurement of various LEC as the experimental tests of variants of the models of the new physics. There are generally two different types of effective lagrangians parametrizing the physics beyond the SM, corresponding to two different scenarios of breaking the gauge symmetry $SU(2)_L\otimes U(1)_Y\rightarrow U(1)_{em}$. The ''decoupling scenario'' assumes that the scale of the new physics is much larger than the electroweak symmetry breaking scale, $% \Lambda \gg v$ , (here $v$ is the vacuum expectation value of the Higgs doublet), the $SU(2)_L\otimes U(1)_Y$ symmetry is then linearly realized and the low energy spectrum is identical with that of the SM, including the Higgs particle, which is supposed to be relatively light. The new nonrenormalizable interactions are organized according to the increasing canonical dimension: \begin{equation} {\cal L}_{eff}={\cal L}_{SM}+\sum_i\frac{f_i^{(6)}}{\Lambda ^2}{\cal O}% _i^{(6)}+\sum_i\frac{f_i^{(8)}}{\Lambda ^4}{\cal O}_i^{(8)}+\ldots \label{linear} \end{equation} and includes all operators of dimension 6, 8, ... invariant with respect to the $SU(2)_L\otimes U(1)_Y$. The ''nondecoupling scenario'' on the other hand corresponds to the case $% \Lambda \approx 4\pi v$. The new physics is related to the symmetry breaking sector of the SM, the gauge symmetry is realized nonlinearly and the effective lagrangian reflects the dynamics of the would-be Goldstone bosons eaten by the gauge bosons (and via the equivalence theorem \cite{ET}, it is related to the dynamics of the longitudinal components of $W^{\pm }$ and $Z$% ). The most economic form of such an effective lagrangian for the symmetry breaking sector is given in the form of the (gauged) nonlinear $\sigma $ model organized as a derivative expansion \begin{equation} {\cal L}_{eff}^{SB}=\frac{v^2}4{\rm tr}(D_\mu U^{+}D^\mu U)+\ldots , \label{chiral} \end{equation} (where $U=\exp ({\rm i}\xi ^a\tau ^a/v)\in SU(2)$, $\tau ^a$ are Pauli matrices and $\xi ^a$ are the would-be Goldstone boson fields, $D_\mu U=\partial _\mu U-g\widehat{W}_\mu U-({\rm i}g^{\prime }/2)B_\mu U$ and $% \widehat{W}_\mu =(1/2{\rm i})W_\mu ^a\tau ^a$); the ellipses here mean terms with four and more derivatives and/or gauge fields, invariant with respect to the local $SU(2)_L\otimes U(1)_Y$ gauge transformation. This has for the $% U$ field the following form \[ U\rightarrow \exp ({\rm i}\alpha _L^a\frac{\tau ^a}2)U\exp (-{\rm i}\alpha _Y% \frac{\tau ^3}2). \] In order to preserve the $\rho $ parameter to be close to one, the additional (global) symmetries are often imposed. The example of such a symmetry is the custodial $SU(2)_c$ symmetry, which is introduced as the unbroken subgroup of the symmetry breaking pattern $SU(2)_L\otimes SU(2)_R\rightarrow SU(2)_c$, completely analogous to the pattern of chiral symmetry breaking of QCD with two light quarks \cite{chpt}. The field $U$ transforms under $SU(2)_L\otimes SU(2)_R$ according to \[ U\rightarrow \exp ({\rm i}\alpha _L^a\frac{\tau ^a}2)U\exp (-{\rm i}\alpha _R^b\frac{\tau ^b}2). \] and the custodial symmetry corresponds to the diagonal subgroup $\alpha _L^a=-\alpha _R^a$. Of course, gauging then the $SU(2)_L\otimes U(1)_Y$ subgroup means the explicit breaking of the $SU(2)_c$ by the terms which vanish for $g^{^{\prime }}\rightarrow 0$. Note also, that there is no Higgs field included in this type of effective lagrangian. However, also in this case it is possible to extend the model to account for such a type of particle \cite{Feruglio} adding to the lagrangian additional terms containing $SU(2)_L\otimes U(1)_Y$ invariant field $H$: \begin{eqnarray} {\cal L}_{eff}^H &=&\frac 12\partial _\mu H\partial ^\mu H-V(H)+a\left( \frac Hv\right) \frac{v^2}4{\rm tr}(D_\mu U^{+}D^\mu U) \nonumber \\ &&\ \ \ +b\left( \frac Hv\right) ^2\frac{v^2}4{\rm tr}(D_\mu U^{+}D^\mu U)+\ldots , \label{nonlinear Higgs} \end{eqnarray} the ellipses stand for terms of higher order as well as for the interaction terms of $H$ field with SM fermions. The SM Higgs is recovered for \begin{eqnarray} V(H) &=&\frac{m_H^2}{8v^2}((H+v)^2-v^2)^2 \nonumber \\ a &=&2,\ b=1. \label{SM Higgs} \end{eqnarray} In the unitary gauge, both these scenarios lead to the $U(1)_{em}$ invariant effective lagrangian with anomalous couplings. E.g. in the gauge boson sector, there are besides other contributions the following triple gauge boson couplings, which are usually written in the form of phenomenological lagrangian (we have omitted possible C or CP violating terms)\footnote{% In fact, in such a phenomenological lagrangian the coupling constant should be understood in general as $q$ dependent formfactors.} \begin{eqnarray} {\cal L}_{eff}^{WWV} &=&\sum_{V=\gamma ,Z}g_V[{\rm i}g_1^V(W_{\mu \nu }^{+}W^\mu V^\nu -W_{\mu \nu }W^{+\mu }V^\nu )+{\rm i}\kappa _VW_\mu ^{+}W_\nu V^{\mu \nu } \\ &&\ \ \ +{\rm i}\frac{\lambda _V}{m_W^2}W_{\nu \mu }^{+}W_\lambda ^\mu V^{\lambda \nu }], \label{TGV} \end{eqnarray} where $W_{\mu \nu }=\partial _\mu W_\nu -\partial _\nu W_\mu $and analogously for $V$, $g_\gamma =-e$ and $g_Z=-g\cos \theta _W$. Note, that within the SM, $g_1^V=\kappa _V=1$, $\lambda _V=0$. The constants $\kappa _\gamma $ and $\lambda _\gamma $ can be interpreted in terms of the anomalous magnetic and quadrupole moment of the $W$ bosons, $g_1^V$ corresponds to the gauge $U(1)$ charges of the $W$ in the the units of $% g_\gamma $ an $g_Z$. Recent constraints on these couplings \cite{experiment} come from the studies of $W\gamma $ events at Fermilab Tevatron; the CDF and D0 results are $-1.6<\Delta \kappa _\gamma <1.8$ and $-0.6<\lambda _\gamma <0.6$. Direct measurement will be also available at LEP2 \cite{LEP2}. \section{Divergences within the effective theory and the dependence on the scale of the new physics} The $U(1)_{em}$ invariant phenomenological lagrangians reviewed in the previous section were often used in various calculations and treated as specific models of the deviations from the SM physics. There was certain controversy in the literature concerning the treatment of the divergences, which appear in the loops with anomalous vertices. Some of the authors used in one or another way the momentum cutoff at the scale of new physics $% \Lambda _{cutoff}=\Lambda _{NP}=\Lambda $. Within this approach, which seems to be very physically illuminating, the principle \cite{Burgess} was used, that the divergent graph cut off at the scale where the effective theory looses its validity due the onset of the new physics gives a lower bound to the actual value of the graph in the full theory. I.e., in practise, only the most divergent contribution of the loop integral to the given amplitude was kept and used as the estimate of the dependence of the full theory amplitude on the scale of the new physics.The appearance of the divergences was therefore interpreted as an indication, that the process under consideration is strongly sensitive to the scale of the new physics. This approach has been criticized (see e.g. the paper \cite{Burgess}) because it is in conflict with the decoupling theorems and gives ambiguous results. Moreover, as usual, it highly overestimates the dependence of the result on the scale\footnote{$\Lambda $ appears in the results calculated within this approach in a positive power (or logarithm) coming from the power counting order of the divergent loops.} $\Lambda $ . However, there are explicit examples discussed e.g. in \cite{Burgess} illustrating that the momentum cutoff could yield correct results\footnote{% The result depends strongly on the choice of the field variables. The correct dependence on the scale of the new physics can be read off from the cutoff dependence of the loops only using the ''good'' variables. However, there is no clear criteria how to decide, whether the choice of variables is ''good'' or ''bad''.}. As usual, it is not possible to resolve whether this is the case before performing explicit calculations. In this paper we would like to briefly illustrate this general situation by using an explicit example of the $W$-boson loops contribution to the process $H\rightarrow \gamma \gamma $ calculated within simple $U(1)_{em}$ invariant phenomenological lagrangian in the unitary gauge using two regularization schemes. We will show, that there are differences between the momentum cutoff approach and another approach based on dimensional regularization with minimal subtractions, which was advocated in \cite{Burgess} and which should be accepted as the procedure giving the correct answer. We will therefore conclude, that the process $H\rightarrow \gamma \gamma $ is not reckoned among the cases, which could be treated correctly within the momentum cutoff prescription in the unitary gauge. The paper is organized as follows. In Section 2 we introduce a specific model of the effective lagrangian with anomalous gauge boson couplings and present the results for the amplitude $H\rightarrow \gamma \gamma $ calculated within the two above mentioned schemes. In Section 3 we discuss the result of the calculations from the general point of view which was sketched in the Introduction. The comparison of the two approaches with respect to the dependence of the decay amplitude on the scale of the new physics and the conclusions are presented in Section 4. The details of the calculation of the amplitude are postponed to the Appendix. \section{$H\rightarrow \gamma \gamma $ within the effective lagrangian approach - a specific model} The theoretical concentration on the decay $H\rightarrow \gamma \gamma $ in the recent literature \cite{Ellis} -- \cite{Hagiwara} is motivated by the fact that this rare decay mode could serve as the main source of the experimental signal for the Higgs particle with the mass within the lower part of the intermediate mass range $m_Z<m_H<2m_W$ on hadron colliders (e.g. LHC). Within the Standard Model, this decay channel of the Higgs boson is described at lowest order by a sum of one-loop Feynman diagrams, this sum is ultraviolet finite (the reason is that there is no tree-level $H\gamma \gamma $ interaction in the SM lagrangian; SM is a renormalizable theory, so that the $H\gamma \gamma $ counterterms cannot be present). As a result, one gets for the decay amplitude an expression of the following form \cite{Ellis}% , the tensor structure of which reflects the Lorentz invariance and $U(1)_{em% {\ }}$gauge invariance: \begin{equation} {\cal M}(H\rightarrow \gamma \gamma )=\frac{e^2g}{16\pi ^2}\frac 1{m_W}\;\varepsilon _\mu ^{\star }\left( k\right) \varepsilon _\nu ^{\star }\left( l\right) \left[ \frac 12m_H^2g^{\mu \nu }-k^\nu l^\mu \right] \;\sum_{i=s,f,g}N_{c_i}e_i^2F_i. \label{SM amplitude} \end{equation} The decay rate is then \begin{eqnarray} \Gamma (H\rightarrow \gamma \gamma )=\frac{\alpha ^3}{128\pi ^2\sin ^2\theta _W}\frac{m_H^3}{m_W^2}\left| \sum_{i=s,f,g}N_{c_i}e_i^2F_i\right| ^2. \label{SM result} \end{eqnarray} Here \begin{eqnarray} F_s &=&\tau _s\left( 1-\tau _sI^2\right) \nonumber \label{eq12} \\ F_f &=&-2\tau _f\left[ 1+\left( 1-\tau _f\right) I^2\right] \nonumber \\ F_g &=&2+3\tau _g+3\tau _g\left( 2-\tau _g\right) I^2 \label{F} \end{eqnarray} are the contributions of the scalar, fermion and vector boson loops resp., \begin{eqnarray} \tau _i &=&4\left( \frac{m_i}{m_H}\right) ^2 \nonumber \label{eq13} \\ N_{c_i} &=&1\;\;\mbox{for}\;i=\mbox{leptons, scalars and vector bosons} \\ N_{c_i} &=&3\;\;\mbox{for}\;i=\mbox{quarks}, \nonumber \end{eqnarray} $e_i$ is electric charge of the loop particle in units of $e$, and \begin{equation} I=\left\{ \matrix{\arctan \frac{1}{\sqrt{\tau-1}} & ,\tau>1 \cr \frac{1}{2} \left[ \pi + i \ln \left| \frac{1+\sqrt{1-\tau}}{1-\sqrt{1-\tau}} \right| \right] & ,\tau<1}\right. . \label{eq14} \end{equation} It was believed that this process could be significantly influenced by possible deviations from the Standard Model, mainly in the gauge boson sector. To parametrize the physics beyond the Standard Model one can employ the effective lagrangian approach sketched above. This was done in several papers (see e.g. \cite{Konig} -- \cite{Hagiwara}); the specific forms of the used effective lagrangians vary among different authors. Since the additional effective couplings are generally of a non-renormalizable type, the resulting decay amplitude is, in contrast to the SM, UV divergent, so that the loop integrals have to be regularized to obtain reasonable predictions. This is another point, at which the above mentioned papers differ. In order to illustrate this general situation we would like to present here a complementary alternative to the treatment contained in the work \cite {Konig}, in which the $W$-boson contribution to the decay width $% H\rightarrow \gamma \gamma $ was calculated within the effective lagrangian approach. Using the same form of an effective lagrangian as in \cite{Konig}, we would like to demonstrate the differences resulting from a different cutoff prescription. We will also shortly comment on the interpretation of the output of the explicit calculations. Let us first briefly review the main results of \cite{Konig} and the way they were obtained. As the phenomenological lagrangian describing the gauge boson sector it was used the $U(1)_{em}$ invariant lagrangian ${\cal {L}}$ introduced originally in \cite{Aronson} (cf. also \cite{Grifols et al} ), conserving $C$ and $P$ separately. In the same notation as in\cite{Konig}, this phenomenological lagrangian is given in the unitary gauge as a sum of three terms \begin{eqnarray} {\cal {L}}={\cal {L}}_1+{\cal {L}}_2+{\cal {L}}_3, \label{lagrangian} \end{eqnarray} where \begin{eqnarray} {\cal {L}}_1 &=&-\frac 12\widehat{G}_{\mu \nu }^{\dagger }\widehat{G}^{\mu \nu }+m_W^2W_\mu ^{\dagger }W^\mu +\frac 12m_Z^2Z_\mu Z^\mu \label{l1} \\ {\cal {L}}_2 &=&-\frac 14\sum_{V=\gamma Z}\widehat{F}_{\mu \nu }^{(V)}% \widehat{F}^{(V)\mu \nu } \label{l2} \\ {\cal {L}}_3 &=&\sum_{V=\gamma Z}\frac{ig_V\lambda _V}{m_W^2}\widehat{F}% _{\mu \nu }^{(V)}\widehat{G}^{\dagger \mu \rho }\widehat{G}_\rho ^{\;\;\nu }. \label{l3} \end{eqnarray} Here \begin{eqnarray} \widehat{G}_{\mu \nu } &=&\left( \partial _\mu -ig_\gamma A_\mu -ig_ZZ_\mu \right) W_\nu -\left( \partial _\nu -ig_\gamma A_\nu -ig_ZZ_\nu \right) W_\mu \nonumber \label{eq5} \\ \widehat{F}_{\mu \nu }^{(V)} &=&F_{\mu \nu }^{(V)}+ig_V\kappa _V\left( W_\mu ^{\dagger }W_\nu -W_\nu ^{\dagger }W_\mu \right) \nonumber \\ F_{\mu \nu }^{(V)} &=&\partial _\mu V_\nu -\partial _\nu V_\mu \nonumber \\ g_\gamma &=&e \nonumber \\ g_Z &=&g\cos \theta _W. \label{structures} \end{eqnarray} Let us note, that the triple gauge boson couplings correspond to (\ref{TGV}) with $g_1^\gamma =g_1^Z=1$. The relation $g_1^\gamma =1$ is quite natural, because it expresses the conservation of the electric charge of $W$ bosons, i.e. the $WW\gamma $ couplings derived from ${\cal {L}}$ are independent linear combination of the most general $U(1)_{em}$ terms conserving $C$ and $% P$. This is, however, not true for the $WW\gamma \gamma $ couplings; here the possible interaction terms are not parametrized by independent LEC but multiplied by very specific combinations of $\kappa _\gamma $ and $\lambda _\gamma .$ For $\kappa _V=1$ and $\lambda _V=0$, the lagrangian ${\cal {L}}$ reduces to the lagrangian of SM in the unitary gauge. In addition to the above lagrangian of the gauge sector, the Higgs boson was included with the standard couplings to $W$ boson pair, \begin{equation} {\cal {L}}_{WWH}=gm_WW_\mu ^{+}W^\mu H. \label{HWW} \end{equation} We can think about the above $U(1)_{em}$ invariant phenomenological lagrangian as being produced by fixing the unitary gauge in the $% SU(2)_L\otimes U(1)_Y$ invariant lagrangian within both decoupling and nondecoupling scenarios. Within the nondecoupling scenario, there are three operators of the order ${\cal O}(p^4)$ in the derivative expansion and three operators of the order ${\cal O}(p^6)$ needed to reproduce the structure of the lagrangian (\ref{lagrangian}) in the unitary gauge, namely\footnote{% In fact, the operator {\rm tr}$(T[V_\mu ,V_\nu ])${\rm tr}$(T[V^\mu ,V^\nu ]) $ can be expressed in terms of the following elements of the operator basis introduced in \cite{Longhitano}: \begin{eqnarray*} {\rm tr}(T[V_\mu ,V_\nu ]){\rm tr}(T[V^\mu ,V^\nu ]) &=&4([{\rm tr}(V_\mu V_\nu )]^2-[{\rm tr}(V_\mu V^\mu )]^2-{\rm tr}(V_\mu V_\nu ){\rm tr}(TV^\mu )% {\rm tr}(TV^\nu ) \\ &+&{\rm tr}(V_\mu V^\mu ){\rm tr}(TV_\nu ){\rm tr}(TV^\nu )). \end{eqnarray*} \par The same can be done with the operator {\rm tr}$(T[V_\nu ,V_\mu ]){\rm tr}(T% \widehat{W}^{\mu \rho }\widehat{W}_\rho ^{\;\;\nu })$, here \par \begin{eqnarray*} {\rm tr}(T[V_\nu ,V_\mu ]){\rm tr}(T\widehat{W}^{\mu \rho }\widehat{W}_\rho ^{\;\;\nu }) &=&{\rm tr}(V_\nu V_\mu ){\rm tr}(\widehat{W}^{\mu \rho }% \widehat{W}_\rho ^{\;\;\nu })-{\rm tr}(V_\mu \widehat{W}^{\mu \rho }){\rm tr}% (\widehat{W}_\rho ^{\;\;\nu }V_\nu ) \\ &&+2{\rm tr}(TV_\mu ){\rm tr}(T\widehat{W}^{\mu \rho }){\rm tr}(\widehat{W}% _\rho ^{\;\;\nu }V_\nu )-2{\rm tr}(T\widehat{W}_\rho ^{\;\;\nu }){\rm tr}% (TV_\mu ){\rm tr}(V_\nu \widehat{W}^{\mu \rho }) \end{eqnarray*} \par \label{footnote}} \begin{eqnarray} {\cal L} &=&\frac 12{\rm tr}\widehat{W}_{\mu \nu }\widehat{W}^{\mu \nu }-\frac 14B_{\mu \nu }B^{\mu \nu }+\frac{v^2}4{\rm tr}(D_\mu U^{+}D^\mu U) \nonumber \\ &&+\frac 1{4g^2}(s_W^2\Delta \kappa _\gamma +c_W^2\Delta \kappa _Z){\rm tr}% (T[V_\mu ,V_\nu ]){\rm tr}(T[V^\mu ,V^\nu ]) \nonumber \\ &&-\frac 1{2g}(s_W^2\Delta \kappa _\gamma +c_W^2\Delta \kappa _Z){\rm tr}(T% \widehat{W}_{\mu \nu }){\rm tr}(T[V^\mu ,V^\nu ]) \nonumber \\ &&+\frac{{\rm i}}{2g}s_Wc_W(\Delta \kappa _\gamma -\Delta \kappa _Z)B_{\mu \nu }{\rm tr}(T[V^\mu ,V^\nu ]) \nonumber \\ &&+\frac 23\frac g{m_W^2}(s_W^2\lambda _\gamma +c_W^2\lambda _Z){\rm tr}(% \widehat{W}_{\nu \mu }\widehat{W}^{\mu \rho }\widehat{W}_\rho ^{\ \;\nu }) \nonumber \\ &&-{\rm i}\frac g{m_W^2}s_Wc_W(\lambda _\gamma -\lambda _Z)B_{\nu \mu }{\rm % tr}(T\widehat{W}^{\mu \rho }\widehat{W}_\rho ^{\;\;\nu }) \nonumber \\ &&-\frac 1{m_W^2}(s_W\lambda _\gamma \Delta \kappa _\gamma +c_W\lambda _Z\Delta \kappa _Z){\rm tr}(T[V_\nu ,V_\mu ]){\rm tr}(T\widehat{W}^{\mu \rho }\widehat{W}_\rho ^{\;\;\nu }), \label{nondecoupling} \end{eqnarray} where \begin{eqnarray*} T &=&U\tau ^3U^{+},\;V_\mu =(D_\mu U)U^{+},\;D_\mu U=\partial _\mu U-g% \widehat{W}_\mu U+g^{^{\prime }}U\widehat{B}_\mu , \\ \widehat{W}_\mu &=&\frac 1{2{\rm i}}W_\mu ^a\tau ^a,\;\widehat{B}_\mu =\frac 1{2{\rm i}}B_\mu \tau ^3,\; \\ \widehat{W}_{\mu \nu } &=&\partial _\mu \widehat{W}_\nu -\partial _\nu \widehat{W}_\mu -g[\widehat{W}_\mu ,\widehat{W}_\nu ],B_{\mu \nu }=\partial _\mu B_\nu -\partial _\nu B_\mu \end{eqnarray*} and $\Delta \kappa _V=\kappa _V-1$. The lagrangian (\ref{HWW}) can be obtained in the same way from ${\cal O}(p^2)$ term \[ {\cal {L}}_{WWH}=\frac{v^2}2\left( \frac Hv\right) {\rm tr}(D_\mu U^{+}D^\mu U). \] Within the decoupling scenario, the same anomalous gauge boson couplings are reproduced by the lagrangian with two dimension 6, three dimension 8 and one dimension 10 operators (cf. e.g. \cite{Renard}) : \begin{eqnarray} {\cal L} &=&\frac 12{\rm tr}\widehat{W}_{\mu \nu }\widehat{W}^{\mu \nu }-\frac 14B_{\mu \nu }B^{\mu \nu }+D_\mu \Phi ^{+}D^\mu \Phi \nonumber \\ &&+\frac{4{\rm i}}{gv^2}s_Wc_W(\Delta \kappa _\gamma -\Delta \kappa _Z)B_{\mu \nu }D^\mu \Phi ^{+}D^\nu \Phi \nonumber \\ &&+\frac 23\frac g{m_W^2}(s_W^2\lambda _\gamma +c_W^2\lambda _Z){\rm tr}(% \widehat{W}_{\nu \mu }\widehat{W}^{\mu \rho }\widehat{W}_\rho ^{\ \;\nu }) \nonumber \\ &&+\frac{16}{gv^4}(s_W^2\Delta \kappa _\gamma +c_W^2\Delta \kappa _Z)\Phi ^{+}\widehat{W}_{\nu \mu }\Phi D^\mu \Phi ^{+}D^\nu \Phi \nonumber \\ &&+\frac 4{gv^4}(s_W^2\Delta \kappa _\gamma +c_W^2\Delta \kappa _Z)[D^\mu \Phi ^{+}D^\nu \Phi -D^\nu \Phi ^{+}D^\nu \Phi ]^2 \nonumber \\ &&+\frac{4{\rm i}}{m_W^2v^2}s_Wc_W(\lambda _\gamma -\lambda _Z)B_{\nu \mu }\Phi ^{+}\widehat{W}^{\mu \rho }\widehat{W}_\rho ^{\ \;\nu }\Phi \nonumber \\ &&+\frac{16}{m_W^2v^4}(s_W\lambda _\gamma \Delta \kappa _\gamma +c_W\lambda _Z\Delta \kappa _Z) \nonumber \\ &&\times (D^\mu \Phi ^{+}D^\nu \Phi -D^\nu \Phi ^{+}D^\nu \Phi )\Phi ^{+}% \widehat{W}^{\mu \rho }\widehat{W}_\rho ^{\ \;\nu }\Phi , \label{decoupling} \end{eqnarray} Here $\Phi $ is the SM Higgs doublet , $s_W=\sin \theta _W,$ $c_W=\cos \theta _W$ and $D^\nu \Phi =\partial _\mu \Phi -g\widehat{W}_\mu \Phi +{\rm i% }g^{^{\prime }}/2B_\mu \Phi .$ The standard Higgs boson coupling (\ref{HWW}) stems now from the third term of (\ref{decoupling}). Such a lagrangian produces, however, also anomalous Higgs boson couplings, which were not considered in \cite{Konig}. Both the nondecoupling and decoupling interpretations of the origin of the phenomenological lagrangian (\ref{lagrangian}) illustrate again the fact, that this lagrangian is incomplete in the sense, that it does not contain all the independent linear combinations of the full set of the operators up to a given dimension or number of (covariant) derivatives\footnote{% Let us also note that the suppression of the higher dimension or higher order operators does not correspond to the usual factors of the type $% 1/\Lambda ^2$ for the decoupling scenario and $1/(4\pi v)$ for the nondecoupling scenario, but rather to the factor $g^2/m_W^2\sim 1/v^2$ . The reason is, that the contributions of all the operators to the parameters $% \kappa _\gamma $ and $\lambda _\gamma $ should have the same order of magnitude in order to reproduce the lagrangian \ref{lagrangian}.}. Nevertheless, we use it here as it stands for the illustrative purposes, mainly because of the relative calculational simplicity. In the paper \cite{Konig}, only the $WWH$, $WW\gamma $ and $WW\gamma \gamma $ vertices derived from the above lagrangian (\ref{lagrangian}) were used in the calculation of the $W$ -boson contribution to the $H\rightarrow \gamma \gamma $ decay width. The calculation was performed in the unitary gauge. There are two types of Feynman diagrams, namely the triangle (and corresponding cross diagram) and the tadpole; both of them, when regulated using momentum cutoff, lead to the quadratic divergent loop integrals. The sum of these diagrams remain UV divergent unless $\kappa _\gamma =1$ and $% \lambda _\gamma =0$, as it was explained above. The result given in \cite {Konig} was presented in the form of the sum of the (cutoff independent) SM expression and the (cutoff dependent) correction. The latter was identified with the quadratic divergence of the diagrams with at least one anomalous vertex proportional $\kappa _\gamma -1$ and/or $\lambda _\gamma .$ I.e. \begin{eqnarray} F_g &=&2+3\tau _g+3\tau _g\left( 2-\tau _g\right) I^2 \nonumber \\ &&\ +\left( \frac \Lambda {m_W}\right) ^2\left( 3\Delta \kappa _\gamma -4\lambda _\gamma +\lambda _\gamma ^2+\frac 12\Delta \kappa _\gamma ^2\right) , \label{FKonig} \end{eqnarray} and the cutoff of the loop momentum $\Lambda $ was interpreted as the scale where the new physics comes in. Because the divergences are local, within this approach the effect of the loops with anomalous gauge boson coupling is equivalent to some direct $H\gamma \gamma $ interaction vertex. It is not difficult to see, that it corresponds to the vertex of the type \[ {\cal L}_{H\gamma \gamma }=G_{eff}e^2\left( \frac Hv\right) F_{\mu \nu }F^{\mu \nu }, \] where the effective coupling constant reads \begin{equation} G_{eff}=\frac 1{(4\pi )^2}\left( \frac \Lambda {m_W}\right) ^2\left( 3\Delta \kappa _\gamma -4\lambda _\gamma +\lambda _\gamma ^2+\frac 12\Delta \kappa _\gamma ^2\right) . \label{treecutoff} \end{equation} As an alternative to this, let us present here the result of our calculation (the details of the calculation can be found in the Appendix) of the same quantity $F_g$ within dimensional regularization and $\overline{MS}$ subtraction scheme. Such a treatment was advocated in \cite{Burgess}. The result can be split up to the finite and divergent parts in the following way \begin{equation} F_g=F_g^{fin}+F_g^{div}, \label{eq14_2} \end{equation} where the finite part is \begin{eqnarray} F_g^{fin} &=&3-{{\kappa _\gamma }^2+3\,\tau _g}+2\,\lambda _\gamma +6\,\kappa _\gamma \,\lambda _\gamma -5\,{{\lambda _\gamma }^2} \nonumber \label{eq15} \\ &&\ {+\biggl( -1+2\,\kappa _\gamma -{{\kappa _\gamma }^2}+2\,\lambda _\gamma -2\,\kappa _\gamma \,\lambda _\gamma -{{\lambda _\gamma }^2}} \nonumber \\ &&\ {+\left( 3+2\,\kappa _\gamma +{{\kappa _\gamma }^2}-2\,\lambda _\gamma -2\,\kappa _\gamma \,\lambda _\gamma +{{\lambda _\gamma }^2}\right) \,\tau _g-3\,{{\tau _g}^2}\biggr) \,{I^2}} \nonumber \\ &&\ +({{-3-2\,\kappa _\gamma +5\,{{\kappa _\gamma }^2}-2\,\lambda _\gamma +2\,\kappa _\gamma \,\lambda _\gamma +\frac{25}3\,{{\lambda _\gamma }^2}}% )\frac 1\tau _g} \nonumber \\ &&\ {{+{{\Bigl(8-4\,\kappa _\gamma -4\,{{\kappa _\gamma }^2}+8\,\lambda _\gamma +8\,\kappa _\gamma \,\lambda _\gamma -14\,{{\lambda _\gamma }^2}}}}} \nonumber \\ &&\ +{{(-4+4\,{{\kappa _\gamma }^2}+8\,{{\lambda _\gamma }^2})}\frac 1\tau _g% } \nonumber \\ &&\ +{\ \left( -4+4\,\kappa _\gamma -8\,\lambda _\gamma -8\,\kappa _\gamma \,\lambda _\gamma +6\,{{\lambda _\gamma }^2}\right) \,\tau _g\Bigr)\,\frac IJ% } \nonumber \\ &&\ +\left( -8\,\lambda _\gamma -4\,\kappa _\gamma \,\lambda _\gamma +6\,{{% \lambda _\gamma }^2}+{{(2-2\,{{\kappa _\gamma }^2}-4\,{{\lambda _\gamma }^2})% }\frac 1\tau _g}\right) \,\ln \left( \frac{m_W{^2}}{{{{\mu }^2}}}\right) \nonumber \\ && \label{Ffinite} \end{eqnarray} and the divergent part is \begin{eqnarray} F_g^{div}{} &{}&{=}\left( -8\,\lambda _\gamma -4\,\kappa _\gamma \,\lambda _\gamma +6\,{{\lambda _\gamma }^2}+{{(2-2\,{{\kappa _\gamma }^2}-4\,{{% \lambda _\gamma }^2})}\frac 1\tau _g}\right) \nonumber \\ &&\ \times {\,\left( {\frac 2{{4-{\rm D}}}}-\gamma _E+\ln (4\,\pi )\right) .} \label{Fdivergent} \end{eqnarray} In these formulae ${\rm D}=4-2\epsilon $ and \begin{equation} J=\left\{ \matrix{ \sqrt{\tau_g-1},& \; \tau_g>1 \cr -i\sqrt{1-\tau_g},& \; \tau_g<1}\right. . \nonumber \end{equation} \section{Discussion} Let us now briefly discuss how to interpret this result. As we have shown above, the lagrangian (\ref{lagrangian}) is a mixture of terms stemming from operators with different dimensions (from 6 up to 10 in the framework of the decoupling scenario) and different orders in the momentum expansion (from $% {\cal O}(p^2)$ up to ${\cal O}(p^6)$ in the framework of the nondecoupling scenario). As a consequence, the formulas (\ref{Ffinite}), (\ref{Fdivergent}% ) do not respect the hierarchy of contributions originating from the hierarchy of the tower of effective operators, which is the cornerstone of the consistent treatment of the nonrenormalizable couplings within the effective lagrangian approach. Therefore, in order to extract the partial information about the relevant dependence on the scale of the new physics $% \Lambda $, it is necessary to reorganize the resulting formulas and to keep only the terms, which are dominant within the two possible scenarios reviewed in the Sec. 1. Within the framework of the decoupling scenario, we expect that we can safely neglect\footnote{% However,as it was shown in \cite{Artz}, in the case when the fundamental full theory is weakly coupled gauge theory, the operators of dimension 8 should also contribute significantly provided they can be generated at the tree level. In this case their contribution is comparable with that of the one loop generated dimension 6 operators, which are multiplied by additional factor $1/16\pi ^2$. This factor can be of the same size like the suppression factor $v^2/\Lambda ^2$ of the dimension 8 operators, provided the scale of new physics is in the range of a few TeV.\label{dimsix}} the operators of dimension 8 and higher and keep only operators of dimension 6. The lagrangian (\ref{decoupling}) contains two such operators (we use here the notation of \cite{HISZ}), namely \begin{eqnarray} {\cal O}_B &=&\frac{{\rm i}g^{^{\prime }}}2B_{\mu \nu }D^\mu \Phi ^{+}D^\nu \Phi , \nonumber \\ {\cal O}_{WWW} &=&-\frac{g^3}{3!}{\rm tr}(\widehat{W}_{\nu \mu }\widehat{W}% ^{\mu \rho }\widehat{W}_\rho ^{\;\;\nu }), \label{OBOWWW} \end{eqnarray} so that we can use (\ref{Ffinite},\ref{Fdivergent}) to get information about the contribution of these two operators only. As far as the $WW\gamma $ and $WW\gamma \gamma $ couplings used for the above calculation of the decay amplitude are concerned, the presence of the term $(\alpha _B/\Lambda ^2)\,{\cal O}_B$ in the lagrangian generates effectively a contribution to the phenomenological parameter $\Delta \kappa _\gamma $ (and not to the $\lambda _\gamma $) \begin{equation} \Delta \kappa _\gamma ^B=\frac 12\frac{m_W^2}{\Lambda ^2}\alpha _B, \label{kappaB} \end{equation} while the term $(\alpha _{WWW}/\Lambda ^2)\,{\cal O}_{WWW}$ contributes to the $\lambda _\gamma $ (and not to the $\Delta \kappa _\gamma $) \begin{equation} \lambda _\gamma ^{WWW}=-\frac 14g^2\frac{m_W^2}{\Lambda ^2}\alpha _{WWW}. \label{lambdaWWW} \end{equation} I.e., within the decoupling scenario, the natural values of the parameters $% \Delta \kappa _\gamma $ and $\lambda _\gamma $ are of the order ${\cal O}(% \frac{m_W^2}{\Lambda ^2})$ and ${\cal O}(g^2\frac{m_W^2}{\Lambda ^2})$ respectively\footnote{% Let us stress that within the full $SU(2)\times U(1)$ approach, the operator ${\cal O}_B$ would induce also the anomalous $HWW\gamma $ coupling; however within the lagrangian (\ref{decoupling}) this contribution is cancelled by the contributions of the higher dimension operators with unnaturally large coupling constants. That is, the lagrangian (\ref{decoupling}) does not allow to get the full information on the influence of the operator ${\cal O}% _B$ on the process $H\rightarrow \gamma \gamma $ using the above formulas (% \ref{Ffinite},\ref{Fdivergent}), which is restricted to the effect of the anomalous $WW\gamma $ and $WW\gamma \gamma $ vertices generated by ${\cal O}% _B$ . On the other hand, the contribution of the loops with one insertion of the anomalous $WW\gamma $ and $WW\gamma \gamma $ vertices generated by dimension 6 operators (\ref{OBOWWW}) can be inferred therefore from (\ref {Ffinite},\ref{Fdivergent}) by means of the expansion to the order ${\cal O}% (\Delta \kappa _\gamma ,\lambda _\gamma ).$} and the leading order anomalous contribution is therefore \begin{eqnarray} F_g^{fin} &=&2+3\tau _g+3\tau _g\left( 2-\tau _g\right) I^2 \nonumber \\ &&\ \ \ \ -2\Delta \kappa _\gamma +8\left( \lambda _\gamma +\Delta \kappa _\gamma \frac 1{\tau _g}\right) -4\lambda _\gamma \tau _gI^2 \nonumber \\ &&\ \ \ \ +4\left[ 4\lambda _\gamma -3\Delta \kappa _\gamma +2\Delta \kappa _\gamma \frac 1{\tau _g}+(\Delta \kappa _\gamma ^B-4\lambda _\gamma )\tau _g\right] \frac IJ \nonumber \\ &&\ \ \ \ -4(3\lambda _\gamma +\Delta \kappa _\gamma ^B\frac 1{\tau _g})\ln \left( \frac{m_W{^2}}{{{{\mu }^2}}}\right) +{\cal O}\left( \frac 1{\Lambda ^4}\right) \label{FOBOWWWfin} \\ F_g^{div} &=&-4\left( 3\lambda _\gamma +\Delta \kappa _\gamma \frac 1{\tau _g}\right) {\left( {\frac 2{{4-{\rm D}}}}-\gamma _E+\ln (4\,\pi )\right) .} \label{FOBOWWWdiv} \end{eqnarray} According to the renormalization prescription for the decoupling scenario, the divergent part should be cancelled by the contributions of the appropriate counterterms stemming from effective operators of dimension 6. In the unitary gauge, such a counterterm has the form\footnote{% Here we explicitly factored out the suppression factor $v^2/\Lambda ^2$ of the dimension 6 operators. There could be also additional suppression of the order $1/(4\pi )^2$ for the one loop generated dimension 6 operators.} \begin{equation} {\cal L}_{H\gamma \gamma }=G_{H\gamma \gamma }\frac{v^2}{\Lambda ^2}% e^2\left( \frac Hv\right) F_{\mu \nu }F^{\mu \nu }, \label{counterterm} \end{equation} where $G_{H\gamma \gamma }$ is an effective (running) coupling constant% \footnote{% This form of interaction is generated e.g. by fixing the unitary gauge in the following dimension 6 operators \par \begin{eqnarray} {\cal O}_{WW} &=&g^2\Phi ^{+}\widehat{W}_{\mu \nu }\widehat{W}^{\mu \nu }\Phi , \nonumber \\ {\cal O}_{BB} &=&-\frac{g^{^{\prime }2}}4\Phi ^{+}B_{\mu \nu }B^{\mu \nu }\Phi , \nonumber \\ {\cal O}_{BW} &=&-\frac{{\rm i}gg^{^{\prime }}}2\Phi ^{+}B_{\mu \nu }% \widehat{W}^{\mu \nu }\Phi . \end{eqnarray} Writing the corresponding terms of the effective lagrangian in the form \par \begin{equation} {\cal L}_{ct}=\frac{\alpha _{WW}}{\Lambda ^2}{\cal O}_{WW}+\frac{\alpha _{BB}% }{\Lambda ^2}{\cal O}_{BB}+\frac{\alpha _{BW}}{\Lambda ^2}{\cal O}_{BW} \label{ctoperators} \end{equation} we have\footnote{% Tree level generated dimension 8 operators could also give significant contribution to $G_{H\gamma \gamma }$, cf. footnote \ref{dimsix} and \cite {mex2}.} \par \begin{equation} G_{H\gamma \gamma }=-\frac 14(\alpha _{WW}+\alpha _{BB}-\alpha _{BW}), \end{equation} Let us also note, that the operators (\ref{ctoperators}) lead to the gauge boson wave function renormalization and mixing, as well as to the anomalous $% HWW\gamma $ and $HWW\gamma \gamma $ vertices, which should also be included in the complete analysis of the process under consideration, however these effects are not discussed here.}. This brings about the following additional contribution to the function $F_g$ : \begin{equation} F_g^{H\gamma \gamma }=\frac{(4\pi v)^2}{\Lambda ^2}G_{H\gamma \gamma }. \label{treedecoupling} \end{equation} . The above result of the loop calculation can be also used to get information on the scale dependence of the effective constant $G_{H\gamma \gamma }.$ In order to ensure the scale independent result for the decay rate it should hold\footnote{% Within the full $SU(2)\times U(1)$ approach, the ellipses here would stand for the other terms coming from the other graphs with vertices generated by $% {\cal O}_B$ as well as from other dimension 6 operators not considered here.} \begin{eqnarray} G_{H\gamma \gamma }(\mu ^{\prime }) &=&G_{H\gamma \gamma }(\mu )-\frac{% \Lambda ^2}{(4\pi v)^2}8\left( 3\lambda _\gamma +\Delta \kappa _\gamma \frac 1{\tau _g}\right) \ln \left( \frac{\mu ^{\prime }}{{\mu }}\right) +\ldots \nonumber \end{eqnarray} Within the framework of the nondecoupling scenario, the lowest order anomalous contribution can be obtained by keeping only the contribution of the ${\cal O}(p^4)$ operators. There are three such operators in the lagrangian (\ref{nondecoupling}). The remaining three operators are of order ${\cal O}(p^6)$, these operators are proportional to the parameters $\lambda _V.$ I.e. setting $\lambda _\gamma \rightarrow 0$ in the formulas (\ref {Ffinite},\ref{Fdivergent}) and expanding to the first order in $\Delta \kappa _\gamma $ we get the leading order contribution to the decay amplitude from the very specific combination of the ${\cal O}(p^4)$ operators (here we use the same notation as in \cite{Longhitano}) \begin{eqnarray} {\cal L}_2 &=&{\rm i}\frac{g^{^{\prime }}}2\ B_{\mu \nu }{\rm tr}(T[V^\mu ,V^\nu ]) \nonumber \\ {\cal L}_4 &=&\ [{\rm tr}(V_\mu V_\nu )]^2 \nonumber \\ {\cal L}_5 &=&\ [{\rm tr}(V_\mu V^\mu )]^2 \nonumber \\ {\cal L}_6 &=&\ {\rm tr}(V_\mu V_\nu ){\rm tr}(TV^\mu ){\rm tr}(TV^\nu ) \nonumber \\ {\cal L}_7 &=&\ {\rm tr}(V_\mu V^\mu ){\rm tr}(TV_\nu ){\rm tr}(TV^\nu )) \nonumber \\ {\cal L}_9 &=&\ {\rm i}\frac g2{\rm tr}(T\widehat{W}_{\mu \nu }){\rm tr}% (T[V^\mu ,V^\nu ]), \label{L2-9} \end{eqnarray} with coefficients given by (\ref{nondecoupling}), cf. also footnote \ref {footnote}. The trilinear vector boson coupling is generated by the operators ${\cal L}_2$ and ${\cal L}_9$. Explicitly, the presence of the terms \begin{equation} {\cal L}=\alpha _2\frac{v^2}{\Lambda ^2}{\cal L}_2+\alpha _9\frac{v^2}{% \Lambda ^2}{\cal L}_9 \end{equation} in the effective lagrangian give rise to the following contribution to the parameter $\Delta \kappa _\gamma $: \begin{equation} \Delta \kappa _\gamma =-g^2\frac{v^2}{\Lambda ^2}(\alpha _2+\alpha _9)=-4% \frac{m_W^2}{\Lambda ^2}(\alpha _2+\alpha _9), \label{kappa29} \end{equation} i.e. the natural value for the coupling $\Delta \kappa _\gamma $ of the lagrangian (\ref{nondecoupling}) is of the order ${\cal O}(\frac{m_W^2}{% \Lambda ^2})$. For the leading order contribution to the function $F_g$ we get then \begin{eqnarray} F_g^{fin} &=&2+3\tau _g+3\tau _g\left( 2-\tau _g\right) I^2 \nonumber \\ &&\ \ \ \ +2\Delta \kappa _\gamma {\biggl( }-1+4\frac 1{\tau _g}+2\left[ -3+2\frac 1{\tau _g}+\tau _g\right] \frac IJ-2\frac 1{\tau _g}\ln \left( \frac{m_W{^2}}{{{{\mu }^2}}}\right) {\biggr) }\ \ \nonumber \\ &&\ \ \ \ +{\cal O}\left( \frac 1{\Lambda ^4}\right) \ \ \end{eqnarray} \begin{equation} F_g^{div}=-4\Delta \kappa _\gamma \frac 1{\tau _g}{\left( {\frac 2{{4-{\rm D}% }}}-\gamma _E+\ln (4\,\pi )\right) .} \label{chiraldiv} \end{equation} Note, that the divergent part is proportional to the $m_H^2=(k+l)^2$, this reflects the fact that in the nondecoupling case the divergencies should be canceled by ${\cal O}(p^6)$ counterterms. In the unitary gauge such a counterterm has the form\footnote{% Here again the suppression factor $v^2m_H^2/\Lambda ^4$ corresponding to the naive dimensional analysis was factored out. The natural value of the symmetry breaking scale is $\Lambda \sim 4\pi v$.} \begin{eqnarray} {\cal L}_{H\gamma \gamma } &=&\widetilde{G}_{H\gamma \gamma }e^2\frac{% v^2m_H^2}{\Lambda ^4}\left( \frac Hv\right) F_{\mu \nu }F^{\mu \nu }, \end{eqnarray} where $\widetilde{G}_{H\gamma \gamma }$ is an effective coupling constant and we have \begin{equation} F_g^{H\gamma \gamma }=\frac{m_H^2}{\Lambda ^2}\widetilde{G}_{H\gamma \gamma }. \label{treenondecoupling} \end{equation} Also in this case we can infer information on the running of this effective coupling; the scale dependence coming from the graphs with the above mentioned specific combination of ${\cal O}(p^4)$ operators should be% \footnote{% Here, within the full nondecoupling approach, the ellipses would mean the contributions of operators not listed above.} \begin{eqnarray} \widetilde{G}_{H\gamma \gamma }(\mu ^{\prime }) &=&\widetilde{G}_{H\gamma \gamma }(\mu )-2\Delta \kappa _\gamma \left( \frac{\Lambda ^2}{4\pi vm_W}% \right) ^2\ln \left( \frac{\mu ^{\prime }}{{\mu }}\right) +\ldots \nonumber \end{eqnarray} \section{Conclusions} Let us now compare these results with those of ref. \cite{Konig}. Inserting the dimensional analysis estimates (\ref{kappaB}, \ref{lambdaWWW}, \ref {kappa29}) of the parameters $\Delta \kappa _\gamma $ and $\lambda _\gamma $ within both scenarios to the cutoff analysis formula and using the principle quoted in the Sec.2 (\ref{FKonig}), we get the following lower bound on the dependence of the function $\Delta F_g$ on the scale of the new physics: \begin{equation} \Delta F_g^{{\rm cutoff}}=\frac 32\alpha _B+g^2\alpha _{WWW}+{\cal O}\left( \frac{m_W^2}{\Lambda ^2}\right) \label{Kdec} \end{equation} for the decoupling scenario and \begin{equation} \Delta F_g^{{\rm cutoff}}=-12(\alpha _2+\alpha _9)+{\cal O}\left( \frac{m_W^2% }{\Lambda ^2}\right) \label{Knon} \end{equation} for the nondecoupling scenario. I.e. in this scheme the natural values of the anomalous contribution are (independently on the scale of the new physics) ${\cal O}\left( 1\right) $ , provided the natural values of the LEC $\alpha _B$, $\alpha _{WWW}$, $\alpha _2$ and $\alpha _9$ are of the order $% {\cal O}\left( 1\right) $. On the other hand, within the dimensional regularization approach the dominant anomalous contribution to the function $% F_g$ comes from the direct interaction terms (cf. formulas (\ref {treedecoupling}, \ref{treenondecoupling})) and the $\mu $ dependent loop logarithms. For the purpose of the dimensional analysis, we can (using the equations for the running of the LEC) interpret the scale $\mu $ as a scale at which the constants $G_{H\gamma \gamma }$and $\widetilde{G}_{H\gamma \gamma }$ acquire their natural values of order ${\cal O}\left( 1\right) $. This is expected to correspond to the point at which the underlying high energy theory is matched with the low energy effective lagrangian, i.e. to the scale of the new physics $\Lambda $. We have then the following estimate \begin{eqnarray*} \Delta F_g^{{\rm DR}} &=&\frac{(4\pi v)^2}{\Lambda ^2}G_{H\gamma \gamma }-\left( \frac 12\frac{m_H^2}{\Lambda ^2}\alpha _B-3g^2\frac{m_W^2}{\Lambda ^2}\alpha _{WWW}\right) \ln \left( \frac{m_W{^2}}{{{{\Lambda }^2}}}\right) +\ldots \\ &=&{\cal O}\left( \frac{(4\pi v)^2}{\Lambda ^2}\right) +{\cal O}\left( \frac{% m_H^2}{\Lambda ^2}\ln \left( \frac{m_W{^2}}{{{{\Lambda }^2}}}\right) \right) +\ldots \end{eqnarray*} for the decoupling case (there could be an overall factor $1/(4\pi )^2$ for the loop generated dimension 6 operator contribution) and \begin{eqnarray*} \Delta F_g^{{\rm DR}} &=&\frac{m_H^2}{\Lambda ^2}\widetilde{G}_{H\gamma \gamma }+4(\alpha _2+\alpha _9)\frac{m_H^2}{\Lambda ^2}\ln \left( \frac{m_W{% ^2}}{{{{\Lambda }^2}}}\right) +\ldots \\ &=&{\cal O}\left( \frac{m_H^2}{\Lambda ^2}\right) +{\cal O}\left( \frac{m_H^2% }{\Lambda ^2}\ln \left( \frac{m_W{^2}}{{{{\Lambda }^2}}}\right) \right) +\ldots \end{eqnarray*} for the nondecoupling case. The ellipses here mean further terms, unimportant from the numerical point of view. We can make the following conclusion.The above formulas show, that the cutoff scheme is in disagreement with dimensional analysis approach in this special case of the calculation of the process $H\rightarrow \gamma \gamma $% . In the often considered cases (e.g. $\Lambda =1TeV$ for the decoupling scenario and loop generated dimension 6 operators, or $\Lambda \leq 4\pi v$ and $m_H<\Lambda $ for the nondecoupling scenario) the formulas (\ref{Kdec}% ), (\ref{Knon}) overestimate the enhancement or suppression of the decay rate of the process under consideration. Therefore, this explicit example does not rank among the cases, which could be treated correctly within the momentum cutoff prescription. This corresponds to the general expectations expressed in the ref. \cite{Burgess}, where further examples of both the failure and success of cutoff analysis within the unitary gauge can be found. \bigskip \noindent{\bf Acknowledgements} We would like to thank to J. Ho\v{r}ej\v{s}\'{\i} for the encouraging discussions and for careful reading of the manuscript and useful remarks on it. This work has been supported in part by research grants GA\v{C}R-1460/95 and GAUK-166/95. \begin{figure}[tbp] \epsfig{file=fig1.eps} \caption{Feynman rules for the $WW\gamma $ and $WW\gamma \gamma $ couplings. All the momenta are in-going. The explicit expressions for the vertex functions are given in the Appendix.} \end{figure} \section*{Appendix} In this Appendix, we display some explicit formulas illustrating the calculation of the decay rate $H\rightarrow \gamma \gamma $ within the framework of the dimensional regularization. The $W$ -boson propagator in the unitary gauge is given by the formula \[ \Delta _F^{\mu \nu }(k)=-\frac{{\rm i}\left( g^{\mu \nu }-\frac{k_\mu k_\nu }{m_W^2}\right) }{k^2-m_W^2+{\rm i}\varepsilon } \] and the Feynman rules for the $HWW$, $WW\gamma $ and $WW\gamma \gamma $ couplings are depicted in Fig. 1 The corresponding vertex functions are \begin{eqnarray} \Gamma _{\alpha \beta \delta }^{(3)}(k,p,r) &=&{\rm i}e\{g_{\alpha \beta }(p-k)_\delta +g_{\alpha \delta }(k-r)_\beta +g_{\beta \delta }(r-p)_\alpha \nonumber \\ &&+\Delta \kappa _\gamma (g_{\beta \delta }r_\alpha -g_{\alpha \delta }r_\beta ) \nonumber \\ &&+\frac{\lambda _\gamma }{m_W^2}[(k\cdot p)(g_{\beta \delta }r_\alpha -g_{\alpha \delta }r_\beta )+(r\cdot k)(g_{\alpha \beta }p_\delta -g_{\beta \delta }p_\alpha ) \nonumber \\ &&+(r\cdot p)(g_{\alpha \delta }k_\beta -g_{\alpha \beta }k_\delta )+r_\beta k_\delta p_\alpha -r_\alpha k_\beta p_\delta ]\} \label{WWG} \end{eqnarray} and \begin{eqnarray} \Gamma _{\alpha \beta \delta \tau }^{(4)}(k,p,r,l) &=&-{\rm i}e^2(2g_{\alpha \beta }g_{\delta \tau }-g_{\alpha \delta }g_{\beta \tau }-g_{\alpha \tau }g_{b\delta }) \\ &&+\frac{{\rm i}e^2\lambda _\gamma }{m_W^2}\{-g_{\alpha \beta }g_{\delta \tau }[(r+l)\cdot (k+p)] \nonumber \\ &&+g_{\alpha \delta }g_{\beta \tau }[(r\cdot p)+(l\cdot k)] \nonumber \\ &&+g_{\alpha \tau }g_{\beta \delta }[(r\cdot k)+(l\cdot p)] \nonumber \\ &&+g_{\alpha \delta }[(k-p)_\tau r_\beta -r_\tau k_\beta -l_\beta k_\tau ] \nonumber \\ &&+g_{\alpha \tau }[(k-p)_\delta l_\beta -r_\beta p_\delta -l_\delta k_\beta ] \\ &&+g_{\beta \delta }[-(k-p)_\tau r_\alpha -r_\tau p_\alpha -l_\alpha p_\tau ] \nonumber \\ &&+g_{\beta \tau }[-(k-p)_\delta l_\alpha -l_\delta p_\alpha -r_\alpha p_\delta ] \nonumber \\ &&+g_{\alpha \beta }[(k+p)_\delta r_\tau +(k+p)_\tau l_\delta ] \nonumber \\ &&+g_{\delta \tau }[k_\beta (r+l)_\alpha +p_\alpha (r+l)_\beta ]\}. \label{WWGG} \end{eqnarray} \begin{figure}[tbp] \epsfig{file=fig2.eps} \caption{The triangle and bubble graphs contributing to one loop amplitude of the decay $H\rightarrow \gamma \gamma $.} \end{figure} In these expressions, the first rows represent the standard model vertices. The decay amplitude is then given by the formula \begin{equation} {\cal M}(H\rightarrow \gamma \gamma )=\varepsilon ^{\star \mu _1}(k_1)\varepsilon ^{\star \mu _2}(k_2){\cal M}_{\mu _1\mu _2}, \label{eq135} \end{equation} where $k_{1,2}$ are the momenta of the out-going photons and $\varepsilon $% 's are their polarization vectors. The polarization tensor ${\cal M}_{\mu _1\mu _2}$ can be splitted into two parts \begin{equation} {\cal M}_{\mu _1\mu _2}={\cal M}_{\mu _1\mu _2}^a+{\cal M}_{\mu _1\mu _2}^b, \label{eq136} \end{equation} where ${\cal M}_{\mu _1\mu _2}^a$ is the contribution of the triangle shown in Fig. 2a (and correspondig crossed graph) and ${\cal M}_{\mu _1\mu _2}^b$ corresponds to the bubble in Fig. 2b, i.e. \begin{eqnarray} {\cal M}_{\mu _1\mu _2}^a &=&\int \frac{{\rm d}^{{\rm D}}l}{(4\pi )^{{\rm D}}% }\frac{{\cal A}_{\mu _1\mu _2}^a(k_{1,}k_2,l)}{\left[ \left( l+k_1\right) ^2-m_W^2\right] \left[ \left( l-k_2\right) ^2-m_W^2\right] \left[ l^2-m_W^2\right] } \nonumber \label{eq137} \\ &&+\int \frac{{\rm d}^{{\rm D}}l}{(4\pi )^{{\rm D}}}\frac{{\cal A}_{\mu _2\mu _1}^a(k_2,k_1,l)}{\left[ \left( l+k_2\right) ^2-m_W^2\right] \left[ \left( l-k_1\right) ^2-m_W^2\right] \left[ l^2-m_W^2\right] } \\ {\cal M}_{\mu _1\mu _2}^b &=&\int \frac{{\rm d}^{{\rm D}}l}{(4\pi )^{{\rm D}}% }\frac{{\cal A}_{\mu _1\mu _2}^b(k_{1,}k_2,l)}{\left[ l^2-m_W^2\right] \left[ \left( l-k_1-k_2\right) ^2-m_W^2\right] }. \end{eqnarray} Here \begin{eqnarray} {\cal A}_{\mu _1\mu _2}^a(k_{1,}k_2,l) &=&{\rm i}gm_W\left( g_{\alpha \mu }-% \frac{(k_1+l)_\alpha (k_1+l)_\mu }{m_W^2}\right) \Gamma _{\mu \nu \mu _1}^{(3)}(k_1+l,-l,-k_1)\left( g_{v\kappa }-\frac{l_\nu l_\kappa }{m_W^2}% \right) \nonumber \\ &&\times \Gamma _{\kappa \lambda \mu _2}^{(3)}(l,-l+k_2,-k_2)\left( g_{\lambda \alpha }-\frac{(l-k_2)_\lambda (l-k_2)_\alpha }{m_W^2}\right) \end{eqnarray} and \begin{eqnarray} {\cal A}_{\mu _1\mu _2}^b(k_{1,}k_2,l) &=&{\rm i}gm_W\left( g_{\alpha \sigma }-\frac{(k_1+k_2+l)_\alpha (k_1+k_2+l)_\sigma }{m_W^2}\right) \nonumber \\ &&\times \Gamma _{\sigma \beta \mu _1\mu _2}^{(4)}(k_1+k_2+l,-l,-k_1,-k_2)\left( g_{\beta \alpha }-\frac{l_\beta l_\alpha }{m_W^2}\right) . \end{eqnarray} Using the standard Feynman parametrization and shifting the loop momenta we get the following reprezentation, which allows for symmetric integration over the loop momentum according to the standard formulas for the $D$% -dimensional integration: \begin{eqnarray} \label{triangle} {\cal M}_{\mu _1\mu _2}^a=2\int_0^1{\rm d}uu{\rm d}v\int \frac{{\rm d}^{{\rm % D}}l}{(4\pi )^{{\rm D}}}\frac{{\cal A}_{\mu _1\mu _2}^a(k_{1,}k_2,l-(uvk_1-u(1-v)k_2))+((k_{1,}\mu _1)\leftrightarrow (k_{2,}\mu _2))}{(l^2-C^a(u,v))^3}, \nonumber \\ \end{eqnarray} where \begin{equation} C^a(u,v)=m_W^2-m_H^2u^2v(1-v) \end{equation} and \begin{equation} {\cal M}_{\mu _1\mu _2}^b=\int_0^1{\rm d}u\int \frac{{\rm d}^{{\rm D}}l}{% (4\pi )^{{\rm D}}}\frac{{\cal A}_{\mu _1\mu _2}^b(k_{1,}k_2,l-u(k_1+k_2))}{% (l^2-C^b(u))^2}, \label{bubble} \end{equation} where \begin{equation} C^b(u)=m_W^2+m_H^2u(1-u). \end{equation} The rest of the calculation ( {\it i.e.} expansion of the numerators of the integrands (\ref{triangle}) and (\ref{bubble}), the symmetric integration over the loop momenta, the integration over the Feynman parameters and extraction of the finite and divergent parts) was performed using {\it \ Mathematica}. We also used the {\it Mathematica }package {\em FeynCalc} \cite {feyncalc}, which proves to be extremely useful for this purpose. \bigskip \bigskip
\section{Introduction} The globular cluster system of the inner Milky Way is still not well understood. This is particularly true for the clusters' classification with respect of the galactic population structure. Because reliably determined parameters, e.g. metallicity, reddening, distance and age, are the basic requirement of any discussion, we present new photometry in (V, I) of the metal-rich globular clusters (GC's) \object{NGC~5927}, 6316, 6342, 6441 and 6760. We also re-discuss \object{NGC~6528} and \object{NGC~6553}, where the data have already been published (Richtler et al. \cite{RTL98}, Sagar et al. \cite{SAG98}). As there has been evidence for a correlation between metallicity and spatial distribution of the GC's since the late 1950's, Zinn (\cite{ZIN85}) classified the clusters via their kinematics, spatial distribution and metallicity into two subsystems: The disk-system with clusters of metallicity $[\mbox{M}/\mbox{H}] \ge -0.8$ dex and the halo-system with $[\mbox{M}/\mbox{H}] \le -0.8$ dex. The disk-system shows a high rotational velocity and a small velocity dispersion, the halo-system vice versa. Armandroff (\cite{ARM89}) derived a scale height of $1.1$ kpc for the disk-system, which he identified with the galactic thick disk via rotational velocities and velocity dispersions. By comparing the metal-rich GC's of the inner $3$ kpc with the underlying stellar population, Minniti (\cite{MIN95}) assigned these objects to the bulge rather than to the disk. Burkert \& Smith (\cite{BUR97}) used kinematical arguments and the masses of the clusters to divide the metal-rich subsystem of Zinn (\cite{ZIN85}) into a bulge, a bar and a disk-group. The thing these subdivisions have in common is, that they refer to the entire system of clusters and they try to formulate their criteria by identifying subsystems within the whole system. For the other way around, i.e. to classify observed objects with any of these subgroups, accurate parameters are needed. The halo clusters are well discernible from any other subsystem, but the metal-rich clusters near the galactic center are not. The determination of their para\-meters encounters observational difficulties, as their low galactic latitudes lead to strong contamination with field stars and to strong (differential) reddening. These effects have to be taken care of. There is a variety of photometry existing for the program clusters. Recent studies on \object{NGC~5927} were done by Fullton et al. (\cite{FUL96}) and Samus et al. (\cite{SAM96}). Armandroff (\cite{ARM88}) presented and discussed CMDs including \object{NGC~6316}, 6342 and 6760. There is a photometry of \object{NGC~6441} in (B,V) of Hesser \& Hartwick (\cite{HES76}) and a more recent one in Rich et al. (\cite{RIC97}). CMDs of \object{NGC~6528} have been discussed by Ortolani et al. (\cite{ORT90}) and Richtler et al. (\cite{RTL98}). Guarnieri et al. (\cite{GUA98}) as well as Sagar et al. (\cite{SAG98}) present (V,I)-photometry of \object{NGC~6553}. Zinn (\cite{ZIN85}), Armandroff (\cite{ARM89}), Richtler et al. (\cite{RTL94}), Minniti (\cite{MIN95}) and Burkert \& Smith (\cite{BUR97}) discuss the subdivision of the GC-system into a halo, disk and/or bulge component. As the data and their reduction shall be published in a forthcoming paper, section 2 deals only briefly with this subject. In section 3 and 4, the derived CMDs are presented and the effects of differential reddening are discussed and removed. Section 5 contains the methods and results of the parameter determination, and in section 6 we will discuss the resulting classification and its problems. \section{Observations and reduction} The observations in V and I were carried out at La Silla/Chile between July 16th and 19th 1993. We used the 2.2m with CCD ESO \#19, which covers with $1024 \times 1024 $ pix an area of $5.7' \times 5.7'$ on the sky. The seeing was $1.1''$. In addition to the ground-based data, we used data of the Hubble-Space-Telescope (HST) for \object{NGC~5927}. These data have already been published by Fullton et al. (\cite{FUL96}). To reduce the data, we used the DAOPHOT-package (Stetson \cite{STE87}, \cite{STE92}), together with the ESO-MIDAS-system (version 1996). The calibrating equations (\ref{equCalib}), determined via Landolt standard stars (Landolt 1992), are \begin{eqnarray} V_{st} = V_{inst} - (1.38 \pm 0.05) + (0.057 \pm 0.003)(V-I)_{st}\nonumber \\ I_{st} = I_{inst} - (2.62 \pm 0.05) - (0.060 \pm 0.003)(V-I)_{st} \label{equCalib} \end{eqnarray} Together with the error of the photometry and of the PSF-aperture-shift, we get an absolute error for a single measurement of $\pm 0.06$ mag. For a more extensive treatment of the data and their reduction, see Heitsch \& Richtler (\cite{HEI99}). To calibrate the HST-data, we used the relations and coefficients as described by Holtzman (\cite{HOL95}). In agreement with Guarnieri et al. (\cite{GUA98}), we detected a systematic shift between calibrated ground-based and HST-magnitudes of about $0.2$ mag with the HST-magnitudes being fainter. This difference might be due to crowding influences on the calibration stars in the ESO-frame, as explained by Guarnieri et al. \section{Colour-Magnitude-Diagrams} \subsection{\object{NGC~5927}} The unselected CMD for \object{NGC~5927} is shown in Fig. \ref{dia5927all}. The cluster's HB and RGB are clearly distinguishable, with the HB overlapping the RGB, as well as the stars of some field population to the blue of the cluster's structures covering the TOP-region. Some $0.5$ mag below the HB, the RGB-bump is discernible. The elongation of the HB and the broadening of the RGB are due to differential reddening, as we now argue. As the HB of metal-rich GCs generally is rather clumped and the HB-stars all have the same luminosity, differential reddening should cause an elongation of the HB parallel to the reddening vector. \begin{figure}[h] \begin{picture}(6.0,3.0) \put(0.0,0.2){\makebox(5.0,3.0){\epsfig{file=./8609f01.eps,scale=0.45 ,bbllx=4.7cm,bblly=4.0cm}}} \end{picture} \caption{Symbols for brightness and colour of the HB-stars in diagrams \ref{dia5927HB} to \ref{dia6760HB}. With the box covering the HB-region of a cluster, all the stars in one of the nine subfields are denoted by the corresponding symbol. Increasing symbol size denotes increasing brightness. Increasing colour-index is represented by a change from squares to lozenges to triangles.\label{diaHBexplic}} \end{figure} \begin{figure*}[h] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f02.ps,scale=0.48 ,bbllx=20.3cm,bblly=18.0cm}}} \end{picture} } \put(4.4,8.2){\makebox(0.6,0.6){\bf N}} \put(8.4,4.4){\makebox(0.4,0.6){\bf E}} \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f03.ps,scale=0.48 ,bbllx=20.3cm,bblly=18.0cm}}} \end{picture} } \put(13.6,8.2){\makebox(0.6,0.6){\bf N}} \put(17.6,4.4){\makebox(0.4,0.6){\bf E}} \end{picture} \hfill \parbox{8.8cm}{ \caption{Coordinates of all the stars in \object{NGC~5927} between $50 \le r \le 400$ pix, r being the distance to the cluster's center in pixel. The HB-stars are marked according to Fig. \ref{diaHBexplic}. Blue bright stars are to be found in an area $0 \le x \le 500$ pix and $250 \le y \le 575$ pix. The stars are selected for photometric errors $\le 0.03$ mag.\label{dia5927HB}}} \hfill \parbox{8.8cm}{ \caption{Coordinates of all the stars in \object{NGC~6342} between $30 \le r \le 250$ pix. The HB-stars are marked according to Fig. \ref{diaHBexplic}. Blue bright stars concentrate in an area to the south of the cluster's center. The stars are selected for photometric errors $\le 0.03$ mag.\label{dia6342HB}}} \end{figure*} \begin{figure*}[h] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f04.ps,scale=0.48 ,bbllx=20.3cm,bblly=18.0cm}}} \end{picture} } \put(4.4,8.2){\makebox(0.6,0.6){\bf N}} \put(8.4,4.4){\makebox(0.4,0.6){\bf E}} \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f05.ps,scale=0.48 ,bbllx=20.3cm,bblly=18.0cm}}} \end{picture} } \put(13.6,8.2){\makebox(0.6,0.6){\bf N}} \put(17.6,4.4){\makebox(0.4,0.6){\bf E}} \end{picture} \hfill \parbox{8.8cm}{ \caption{Coordinates of all the stars in \object{NGC~6441} between $30 \le r \le 400$ pix. The HB-stars are marked according to Fig. \ref{diaHBexplic}. The lower star density for $y \le 360$ pix is due to a coordinate shift of the calibration frames. The stars are selected for photometric errors $\le 0.03$ mag.\label{dia6441HB}}} \hfill \parbox{8.8cm}{ \caption{Coordinates of all the stars in \object{NGC~6760} between $40 \le r \le 350$ pix. The HB-stars are marked according to Fig. \ref{diaHBexplic}. The blue bright stars are found to the south of the cluster's center. The stars are selected for photometric errors $\le 0.03$ mag.\label{dia6760HB}}} \end{figure*} Fig. \ref{dia5927HB} shows the coordinates of the radially selected cluster stars with special markings for the HB-stars as given by Fig. \ref{diaHBexplic}. If differential reddening is indeed responsible for the observed elongation, we do not expect to find any red faint stars in areas where blue bright stars are to be found, unless the reddening is very (!) patchy. In the case of \object{NGC~5927}, we note (Fig. \ref{dia5927HB}) that the blue bright stars are located in an area west of the cluster's center, and thus differential reddening is indeed responsible for the elongated HB structure. In Fig. \ref{dia5927HST} we present the calibrated HST-CMDs of \object{NGC~5927}. They all show a slightly broadened lower RGB as well as a slightly tilted HB. The TOP is well resolved. However, the CMDs do not extend to the bright stars of the AGB/RGB due to pixel overflow on the exposures. As mentioned above, the HST-CMDs are shifted with respect to the ESO-CMDs to fainter magnitudes. Richtler et al. (\cite{RTL98}) argue that the ground-based calibration is not erroneous. Thus, we will use the ground-based calibrationed data for the further analysis. For a detailed discussion see Heitsch \& Richtler (\cite{HEI99}). \subsection{\object{NGC~6316}} The unselected CMD (Fig. \ref{dia6316all}) shows beside the cluster a strong contribution from the field population. The field main sequence is striking. The cluster RGB is broadened, but since the clumpy HB indicates only small differential reddening, the RGB width is probably to a large part due to the field contamination. Determining a correlation between HB-stars and coordinates as in Fig. \ref{dia5927HB} led to no convincing results, because the field does not contain enough stars. \subsection{\object{NGC~6342}} \object{NGC~6342} (Fig. \ref{dia6342all}) shows a sparsely populated AGB/RGB due to the small size of the cluster. The TOP-region and upper MS are reached. Fig. \ref{dia6342HB} gives the location of HB-stars as in Fig. \ref{dia5927HB}. Blue, bright HB-stars are to be found in an area to the south of the cluster's center. \subsection{\object{NGC~6441}} \object{NGC~6441} (Fig. \ref{dia6441all}) is located behind a dense field population, the stars of which can be found between $1.0 \le V-I \le 1.5$ mag. This population covers the lower part of the RGB of \object{NGC~6441} as well. Its TOP is not reached. We mention some special features. First, we find some stars between $1.8 \le V-I \le 2.0$ mag and above the HB of \object{NGC~6441}. These could be HB-stars of a population which is similar to \object{NGC~6441} and is located between the cluster and ourselves, as the stars are shifted in V only. In this case, we would have to assume that the absolute reddening is caused by some cloud between this population and the observer. Otherwise it would have to be shifted in $V-I$ as well. Second, we find some stars to the blue of the clumpy, tilted HB of \object{NGC~6441}. These stars seemingly belong to the cluster, as they are still visible when selecting for small radii. Probably they belong to the blue HB of \object{NGC~6441}, which has been discovered by Rich et al. (\cite{RIC97}). The difference in star density (Fig. \ref{dia6441HB}) is due to the fact that the calibration exposures were shifted by around 360 pix to the south. Taking this into account, we find that the blue bright stars are mostly found in the western two thirds of the cluster. \subsection{\object{NGC~6760}} Fig. \ref{dia6760all} not only shows the already discussed structures such as HB, RGB and field population, but it shows the RGB-bump below the HB as well. The RGB is rather broadened. The HB and RGB-bumps are elongated and tilted with the same slope. The HB-stars, marked according to colour and brightness, are found in Fig. \ref{dia6760HB}. \subsection{\object{NGC~6528} and \object{NGC~6553}} The CMDs for \object{NGC~6528} and 6553 (Fig. \ref{dia6528all}, \ref{dia6553all}) have already been published (Richtler et al. \cite{RTL98}, Sagar et al. \cite{SAG98}). As described in their papers, \object{NGC~6528} not only shows a broadened RGB and a tilted and elongated HB, but it shows also some background population below the AGB/RGB. The field population covers the TOP-region of the cluster. Moreover, the RGB-bump of \object{NGC~6528} is clearly visible some $0.5$ mag below the HB. The CMD of \object{NGC~6553} shows the same characteristics as \object{NGC~6528}, however they are even more distinct. Here we clearly see the background population with its RGB and AGB/RGB strongly differentially reddened. \begin{figure*}[hp] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f06.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f07.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{Unselected CMD for \object{NGC~5927}. RGB and HB are clearly visible as well as some stars of the field population.\label{dia5927all}}} \hfill \parbox{8.8cm}{ \caption{Differentially dereddened CMD for \object{NGC~5927}. The RGB-bump is well discernible now. \label{dia5927dc}}} \end{figure*} \begin{figure*}[hp] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f08.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f09.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{Unselected CMD for \object{NGC~6316}. The AGB/RGB of the field population is well discernible. \label{dia6316all}}} \hfill \parbox{8.8cm}{ \caption{Differentially dereddened CMD for NGC 6316. Due to the only slight differential reddening, the effect of the correction is not as distinct as in \object{NGC~5927}, for example. Moreover, the dereddened CMD contains stars of the inner $2.24'\times 2.24'$ around the cluster's center only (See extinction maps in paragraph \ref{ssecCorDifRed}).\label{dia6316dc}}} \end{figure*} \begin{figure*}[hp] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f10.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f11.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{Unselected CMD for \object{NGC~6342}. TOP and upper MS are reached. \label{dia6342all}}} \hfill \parbox{8.8cm}{ \caption{Differentially dereddened CMD for \object{NGC~6342}. The effect of the correction is best seen at the TOP-region. \label{dia6342dc}}} \end{figure*} \begin{figure*}[hp] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f12.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f13.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{Unselected CMD for \object{NGC~6441}. Strong contamination by the field population.\label{dia6441all}}} \hfill \parbox{8.8cm}{ \caption{Differentially dereddened CMD for \object{NGC~6441}. The effect of the correction shows best in the narrower lower RGB. The HB still partly overlaps with the RGB, which means, that the correction was not completely succesful. This is mostly due to the strong field population. \label{dia6441dc}}} \end{figure*} \begin{figure*}[hp] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f14.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f15.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{Unselected CMD for \object{NGC~6760}. The effect of differential reddening is clearly visible. \label{dia6760all}}} \hfill \parbox{8.8cm}{ \caption{Differentially dereddened CMD for \object{NGC~6760}. The HB lies well to the blue of the RGB, the RGB-bump clearly on the RGB below the HB. \label{dia6760dc}}} \end{figure*} \begin{figure*}[hp] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f16.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f17.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{Unselected CMD for \object{NGC~6528}. Below the AGB/RGB there are traces of the background population. \label{dia6528all}}} \hfill \parbox{8.8cm}{ \caption{Differentially dereddened CMD for \object{NGC~6528}. The HB still overlaps strongly with the RGB. Thus, the correction was not completely successful due to the strong contamination by field stars. However, the RGB narrowed perceptibly, and the RGB-bump is now visible below the HB. \label{dia6528dc}}} \end{figure*} \begin{figure*}[hp] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f18.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f19.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{Unselected CMD for \object{NGC~6553}. Strong contamination by the field population is visible. \label{dia6553all}}} \hfill \parbox{8.8cm}{ \caption{Differentially dereddened CMD for \object{NGC~6553}. RGB and RGB-bump are clearly defined now, and the HB lies well to the blue of the RGB. \label{dia6553dc}}} \end{figure*} \begin{figure*}[hp] \begin{picture}(17.6,9.3) \put(0.0,4.8) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(4.0,4.0){\epsfig{file=./8609f20a.ps,scale=0.24 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(4.6,4.8) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(4.0,4.0){\epsfig{file=./8609f20b.ps,scale=0.24 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(0.0,0.0) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(4.0,4.0){\epsfig{file=./8609f20c.ps,scale=0.24 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(4.6,0.0) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(4.0,4.0){\epsfig{file=./8609f20d.ps,scale=0.24 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(8.0,8.0){\epsfig{file=./8609f21.ps,scale=0.43 ,bbllx=21.0cm,bblly=15.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm} { \caption{\object{NGC~5927}: Unselected HST-CMDs (PC, WF2, WF3 and WF4). The TOP is well resolved. \label{dia5927HST}} } \hfill \parbox{8.8cm} { \caption{Slopes of the HBs, i.e. the reddening vectors against galactic longitude (here in cartesian coordinates).\label{diaExtVar}} } \end{figure*} \section{Correction for differential reddening\label{ssecCorDifRed}} In order to correct the CMDs for differential reddening, we used a refined version of the method described by Grebel et al. (\cite{GRE95}). The entire frame is divided into subframes. These are determined by covering the whole frame with a regular subgrid and dividing the grid cells further until the number of stars in one cell becomes too small to define the CMD structure. The CMDs will be shifted according to the reddening vector (described below) with respect to CMDs from neighbouring cells. The shift in colour supplies the differential reddening. If two neighbouring subframes have the same reddening, these subframes are merged. There are two problems with this method: First, one has to be careful to use the HB as a means for comparing two CMDs, as the HB may be intrinsically elongated. Useful results can only be achieved by comparing the RGBs and TOPs, as far as they are accessible. Second, the size of the subfields must be large enough to render meaningful CMDs. Fig. \ref{diaAllClusExt} shows the resulting extinction maps for the seven clusters. The smallest subfields have a size of about $28'' \times 28''$. But as some of them still showed differential reddening, the scale of the structures responsible for the differential reddening is be expected to be even smaller. The smallest scales we got from a comparison of coordinates of stars with different reddening amounted to values of $4''$. \begin{figure*}[h] \begin{picture}(18.0,19.0) \put(0.0,12.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(5.5,5.5){\epsfig{file=./8609f22a.eps,scale=0.52 ,bbllx=10.5cm,bblly=5.3cm}}} \end{picture} } \put(0.5,17.2){\makebox(1.,0.4){\footnotesize \object{NGC~5927}}} \put(2.5,17.2){\makebox(0.6,0.6){\bf N}} \put(5.1,14.7){\makebox(0.4,0.6){\bf E}} \put(5.8,12.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(5.5,5.5){\epsfig{file=./8609f22b.eps,scale=0.52 ,bbllx=10.5cm,bblly=5.3cm}}} \end{picture} } \put(6.3,17.2){\makebox(1.0,0.4){\footnotesize \object{NGC~6316}}} \put(8.3,17.2){\makebox(0.6,0.6){\bf N}} \put(10.9,14.7){\makebox(0.4,0.6){\bf E}} \put(11.7,12.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(5.5,5.5){\epsfig{file=./8609f22c.eps,scale=0.52 ,bbllx=10.5cm,bblly=5.3cm}}} \end{picture} } \put(12.2,17.2){\makebox(1.0,0.4){\footnotesize \object{NGC~6342}}} \put(14.2,17.2){\makebox(0.6,0.6){\bf N}} \put(16.8,14.7){\makebox(0.4,0.6){\bf E}} \put(0.0,6.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(5.5,5.5){\epsfig{file=./8609f22d.eps,scale=0.52 ,bbllx=10.5cm,bblly=5.3cm}}} \end{picture} } \put(0.5,11.2){\makebox(1.0,0.4){\footnotesize \object{NGC~6441}}} \put(2.5,11.2){\makebox(0.6,0.6){\bf N}} \put(5.1,8.7){\makebox(0.4,0.6){\bf E}} \put(5.8,6.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(5.5,5.5){\epsfig{file=./8609f22e.eps,scale=0.52 ,bbllx=10.5cm,bblly=5.3cm}}} \end{picture} } \put(6.3,11.2){\makebox(1.0,0.4){\footnotesize \object{NGC~6760}}} \put(8.3,11.2){\makebox(0.6,0.6){\bf N}} \put(10.9,8.7){\makebox(0.4,0.6){\bf E}} \put(11.7,6.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(5.5,5.5){\epsfig{file=./8609f22f.eps,scale=0.52 ,bbllx=10.5cm,bblly=5.3cm}}} \end{picture} } \put(12.2,11.2){\makebox(1.0,0.4){\footnotesize \object{NGC~6528}}} \put(14.2,11.2){\makebox(0.6,0.6){\bf N}} \put(16.8,8.7){\makebox(0.4,0.6){\bf E}} \put(5.8,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(5.5,5.5){\epsfig{file=./8609f22g.eps,scale=0.52 ,bbllx=10.5cm,bblly=5.3cm}}} \end{picture} } \put(6.3,5.2){\makebox(1.0,0.4){\footnotesize \object{NGC~6553}}} \put(8.3,5.2){\makebox(0.6,0.6){\bf N}} \put(10.9,2.7){\makebox(0.4,0.6){\bf E}} \end{picture} \caption{Extinction maps for \object{NGC~5927} to \object{NGC~6553}, derived from plotting CMDs for each of the areas. The numbers in the subfields give the absolute reddening derived via isochrone fitting. The differential reddening was determined using the minimal absolute reddening as a point of reference. As the derived scale of reddening variation strongly depends on the number of stars, the resulting scales can only be estimates for upper boundaries. The maps cover $5.7' \times 5.7'$, i.e. they cover the whole area of the original frames. An exception is made with NGC 6136, where we constrained the map to the inner $2.24' \times 2.24'$, because of the clusters small size. \label{diaAllClusExt}} \end{figure*} For the correction of the CMDs we need the extinction \begin{equation} A_V=R^B_V E_{B-V}=R^I_V E_{V-I}\mbox{.} \label{equExtLaw} \end{equation} However, assuming a uniform reddening law led to CMDs which in some cases showed the corrected HBs having larger or smaller slopes than the uncorrected HBs. Moreover, as there is some uncertainty in the literature regarding $R^B_V$, with values varying between $R^B_V=3.1$ (Savage \& Mathis \cite{SAV79}) and $R^B_V=3.6$ (Grebel \& Roberts \cite{GRR95}), we determined the slope of the reddening vector via the tilted HBs of our CMDs. This leads to reasonable results only if the HBs are intrinsically clumpy. This assumption is corroborated by the fact that the well dereddened CMDs (Fig. \ref{dia5927dciso}, \ref{dia6760dciso}, \ref{dia6553dciso}) have clumpy HBs indeed. Table \ref{tabDifRed} shows the slopes $R^I_V$ for each cluster. In Fig. \ref{diaExtVar}, the slopes are plotted against the galactic longitude. These variations, although at the margin of the errors, confirm earlier observations by Meyer \& Savage (\cite{MEY81}) and Turner (\cite{TUR94}). Meyer \& Savage determined via two-color-diagrams the deviation of single stars in the extinction behaviour from the galactic mean extinction law. Turner demonstrated the inapplicability of a mean galactic reddening law for objects lying close to the galactic plane. \begin{table}[h] \footnotesize \begin{center} \begin{tabular}{lcc} \hline NGC & $\Delta E_{V-I}^{max}$ & $R_V^I$ \\ \hline 5927 & $0.27$ & $1.9\pm 0.2$ \\ 6316 & $0.07$ & $2.1\pm 0.2$ \\ 6342 & $0.32$ & $2.2\pm 0.1$ \\ 6441 & $0.20$ & $2.3\pm 0.1$ \\ 6760 & $0.25$ & $2.0\pm 0.1$ \\ 6528 & $0.31$ & $2.4\pm 0.3$ \\ 6553 & $0.29$ & $2.3\pm 0.2$ \\ \hline \end{tabular} \end{center} \normalsize \caption{Maximum differential reddening and slope of the reddening vector. \label{tabDifRed}} \end{table} To correct the diagrams for differential reddening, we referred all sub-CMDs of one cluster to the one with detected minimal reddening and we shifted all other sub-CMDs onto that. As we thus use the minimal absolute reddening as a point of reference, the absolute reddening determined later on will be smaller than value given in the literature. The differentially dereddened CMDs are shown in Figs. \ref{dia5927dc} through \ref{dia6553dc}. As the correction led to a clearly improved appearance for all clusters, the corrected versions of the CMDs will be used for further investigation. \begin{figure*}[hp] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f23.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f24.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{Isochrone fitting for \object{NGC~5927}. $[\mbox{M}/\mbox{H}]=-0.40$ dex, $t=14.5$ gyr. The CMD is selected for photometric errors $\le 0.04$ mag and for radii $(50 \le r \le 400)$ pix. The first selection gives preference to brighter stars, i.e. to stars which not necessarily are cluster members. This effect is counterbalanced by the second selection. \label{dia5927dciso}}} \hfill \parbox{8.8cm}{ \caption{\object{NGC~6316}, selected for photometric errors $\le 0.04$ mag and with an isochrone $[\mbox{M}/\mbox{H}]=-0.70$, $t=14.5$ gyr. The radial selection is due to the correction for differential reddening (see Figs. \ref{dia6316dc} and \ref{diaAllClusExt}). \label{dia6316dciso}}} \end{figure*} \begin{figure*}[h] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f25.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f26.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{Isochrone for \object{NGC~6342} ($[\mbox{M}/\mbox{H}]=-0.40$ dex, $t=14.5$ gyr). The CMD is selected for radii $\le 200$ pix. \label{dia6342dciso}}} \hfill \parbox{8.8cm}{ \caption{CMD of \object{NGC~6441} with isochrone $[\mbox{M}/\mbox{H}]=-0.40$ dex, $t=14.5$ gyr. The CMD is radially selected $(50 \le r \le 400)$ pix. The stars to the blue of the HB still are visible. \label{dia6441dciso}}} \end{figure*} \begin{figure*}[h] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f27.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \put(9.7,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f28.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{The isochrone for \object{NGC~6760} with $[\mbox{M}/\mbox{H}]=-0.40$ dex and $t=14.5$ gyr is slightly too metal-rich, as the AGB/RGB arcs above the model. This result is corroborated by the metallicity- estimates given below. The CMD is radially selected for $100 \le r \le 300$ pix. \label{dia6760dciso}}} \hfill \parbox{8.8cm}{ \caption{The radially selected CMD of \object{NGC~6528} clearly shows the two AGB/RGB's of the cluster's and the background population. The isochrone ($[\mbox{M}/\mbox{H}]=0.0$ dex, $t=14.5$ gyr) is slightly too metal-rich as well. \label{dia6528dciso}}} \end{figure*} \begin{figure*}[h] \begin{picture}(18.0,8.8) \put(0.5,0.2) { \begin{picture}(0.0,0.0) \put(0.0,0.0){\makebox(7.0,7.0){\epsfig{file=./8609f29.ps,scale=0.48 ,bbllx=20.5cm,bblly=18.0cm}}} \end{picture} } \end{picture} \hfill \parbox{8.8cm}{ \caption{Radially selected CMD ($50 \le r \le 400$ pix) of \object{NGC~6553} with an isochrone of $[\mbox{M}/\mbox{H}]=0.0$ dex and $t=14.5$ gyr. The strongly differentially reddened field population is now clearly separated from the cluster (to the red of the lower RGB). The CMD is also selected for photometrical errors $\le 0.05$ mag. \label{dia6553dciso}}} \end{figure*} \clearpage \section{The Globular Cluster Parameters} This section deals with the determination of the metallicity, reddening and distance using the {\it differentially dereddened} CMDs. There are two possible ways to achieve the goal. In the first, theoretical models are compared with the CMDs, in the second, empirical relations between parameters and loci in the CMDs are used. \subsection{Isochrone fitting} To derive metallicity, distance and absolute reddening via isochrone fitting, we used the Padova-tracks (Bertelli et al. \cite{BER94}) with a fixed age of $14.5$ Gyr ($\log(age)=10.160$). Isochrones with different ages ($10.120 \le \log(age) \le 10.200$) led to identical results. To avoid systematic errors, we used the middle of the broadened structures to fit the isochrones by eye. These loci are easily determined for the ascending part of the RGB, as it runs more or less perpendicular to the reddening vector. Regarding the upper part of the RGB, we take into account that we cannot distinguish between the AGB and the RGB in our diagrams. Hence, the densest regions of the AGB/RGB lie between the model's tracks. We additionally used the HB and the lower part of the RGB, as far as they were accessible. The parameters resulting from the isochrone fit are given in Table \ref{tabIso}. \begin{table}[h] \footnotesize \begin{center} \begin{tabular}{lccc} \hline NGC & $(m-M)_V$ & $E_{V-I}$ & $[\mbox{M}/\mbox{H}]$\\ \hline 5927 & $15.45\pm 0.03$ & $0.43\pm 0.02$ & $-0.40$ \\ 6316 & $16.76\pm 0.04$ & $0.62\pm 0.03$ & $-0.70$ \\ 6342 & $15.36\pm 0.04$ & $0.46\pm 0.03$ & $-0.40$ \\ 6441 & $16.48\pm 0.05$ & $0.49\pm 0.03$ & $-0.40$ \\ 6760 & $16.18\pm 0.04$ & $0.72\pm 0.03$ & $-0.40$ \\ 6528 & $15.94\pm 0.05$ & $0.46\pm 0.03$ & $0.00$ \\ 6553 & $15.42\pm 0.04$ & $0.76\pm 0.03$ & $0.00$ \\ \hline \end{tabular} \end{center} \normalsize \caption{Distance modulus $(m-M)_V$, total reddening $E_{V-I}$ and metallicity $[\mbox{M}/\mbox{H}]$ of all the sample's clusters derived by isochrone fitting. The errors are eye-estimates of how accurately we could place the isochrones. Note that the isochrones are fitted to the differentially dereddened CMDs.\label{tabIso}} \end{table} Figures \ref{dia5927dciso} to \ref{dia6553dciso} show the differentially dereddened CMDs with the fitted isochrones. For a discussion and comparison of these parameters with the literature, see paragraph \ref{ssecComp}. \subsection{Metallicity and reddening: relations } \subsubsection{Metallicity} The luminosity difference between HB and the turn over of the AGB/RGB in $(V,V-I)$-CMDs is very sensitive to metallicity in the metal-rich domain (e.g. Ortolani et al. \cite{ORT97}). Moreover, it is a differential metallicity indicator, thus it is independent of absolute colour or luminosity, in contrast to the $[\mbox{M}/\mbox{H}]-(V-I)_{0,g}$-method (see e.g. Sarajedini \cite{SAR94}). We present a preliminary linear calibration of this method, \begin{equation} [\mbox{M}/\mbox{H}]=a(V_{HB}-V_{RGB}^{max}) + b, \label{equFeHdV} \end{equation} as there has not been any so far. Because there still are only very few $(V,V-I)$-CMDs which clearly show both the HB and turn over of the AGB/RGB and which have reliable metallicity determinations, we used the Padova-isochrones and a CMD of NGC 6791 (Garnavich et al. \cite{GAR94}) to set up a calibration. NGC 6791 is one of the richest old open clusters with a good metallicity determination and it is therefore suitable to serve as a zero-point check. As the form of the RGB depends slightly on age as well (e.g. Stetson et al. \cite{STE96}), we have to check this dependence before applying our calibration. Fig. \ref{diaCalibdV} shows the linear relation between $[\mbox{M}/\mbox{H}]$ and $\Delta V \equiv V_{HB} - V_{max}$ for four GC-ages. Table \ref{tabCalibdV} contains the respective coefficients. As the metal-poorest isochrones of the Padova-sample ($[\mbox{M}/\mbox{H}] = -1.70, 1.30$ dex) do not show a maximum of the AGB/RGB, they have not been used. Fig. \ref{diaCalibdV} makes clear that the age has only a minor influence on the resulting metallicity. To be consistent with the isochrone-fit, we used the relation for $\log(age)=10.160$. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{8609f30.ps}} \caption{$[\mbox{M}/\mbox{H}]-\Delta V$-relation for different ages. The open symbols stand for the isochrone-values. Squares, lines: $9.8$ Gyr, diamonds, dotted: $13.2$ Gyr, triangles, dashed: $14.5$ Gyr, circles, dash-dotted: $15.8$ Gyr. The filled square represents the value for NGC 6791 (Garnavich et al. \cite{GAR94}). \label{diaCalibdV}} \end{figure} To estimate the metallicities of our clusters, we now only have to measure the relevant luminosities. The results are given in Table \ref{tabFeHdV}. The value for \object{NGC~5927} given in column $[M/H]_2$ relates to a single star (Fran\c{c}ois \cite{FRA91}); \object{NGC~6528} and \object{NGC~6553} are from Richtler et al. (\cite{RTL98}) and Sagar et al. (\cite{SAG98}). \begin{table} \begin{center} \begin{tabular}{lcc} \hline $\log(age)$ & a [dex/mag] & b [dex] \\ \hline $9.990$ & $-0.614 \pm 0.020$ & $0.929 \pm 0.026$ \\ $10.120$ & $-0.595 \pm 0.017$ & $0.805 \pm 0.024$ \\ $10.160$ & $-0.602 \pm 0.032$ & $0.800 \pm 0.043$ \\ $10.200$ & $-0.592 \pm 0.033$ & $0.764 \pm 0.045$ \\ \hline \end{tabular} \end{center} \caption{Calibration coefficients of the $[M/H]-\Delta V$-relation for different cluster ages. The calibration equation is $[M/H]$ $=a\Delta V + b$. \label{tabCalibdV}} \end{table} \begin{table*} \begin{center} \begin{tabular}{lcccccc} \hline NGC &$V_{HB}$ &$V_{RGB}^{max}$&$\Delta V$&$[M/H]$[dex] &$[M/H]_1$&$[M/H]_2$\\\hline 5927&$16.30\pm 0.03$&$14.49\pm 0.04$&$1.81$ &$-0.29\pm 0.08$&$-0.37$ &$-1.08$ \\ 6316&$17.42\pm 0.05$&$15.09\pm 0.04$&$2.33$ &$-0.60\pm 0.09$&$-0.55$ & \\ 6342&$16.47\pm 0.05$&$14.44\pm 0.06$&$2.03$ &$-0.42\pm 0.09$&$-0.65$ & \\ 6441&$17.51\pm 0.07$&$15.53\pm 0.07$&$1.98$ &$-0.39\pm 0.10$&$-0.53$ & \\ 6760&$17.08\pm 0.02$&$14.81\pm 0.04$&$2.27$ &$-0.57\pm 0.09$&$-0.52$ & \\ 6528&$17.07\pm 0.06$&$15.74\pm 0.08$&$1.33$ &$ 0.00\pm 0.09$&$-0.17$ &$-0.15$ \\ 6553&$16.52\pm 0.07$&$15.13\pm 0.06$&$1.39$ &$-0.04\pm 0.08$&$-0.25$ &$-0.10$ \\ \hline \end{tabular} \end{center} \caption{Metallicities of all GCs via the differentially dereddened CMDs. Column $[M/H]$ contains the values derived by the $[M/H]-\Delta V$-relation. The values of column $[\mbox{M}/\mbox{H}]_1$ have been taken from Harris (\cite{HAR96}). Column $[M/H]_2$ gives additional values as discussed in the text. The errors only take account of the uncertainties of the luminosities and the calibration errors of Table \ref{tabCalibdV}. \label{tabFeHdV}} \end{table*} \subsubsection{Reddening} It should be remembered, that we used the differentially dereddened CMDs to determine the parameters. Thus, the given reddenings are minimal ones. As mentioned above, the absolute colour of the RGB at the level of the HB can be used to estimate the metallicity. Conversely (Armandroff \cite{ARM88}), if we know the metallicity, we can determine the absolute colour $(V-I)_{0,g}$ and thus the absolute reddening of the cluster. These relations between colour $(V-I)_{0,g}$ and metallicity are well calibrated for the metal-poor to intermediate regime. However, it is difficult to set up a calibration for the metal-rich regime of our clusters. Linear calibrations have been provided by e.g. Sarajedini (\cite{SAR94}). A more recent calibration by Caretta \& Bragaglia (\cite{CAA98}) uses a 2nd order polynomial. To set up a calibration for the metal-rich regime we used again the Padova-tracks together with NGC 6791 to derive the coefficients for a relation of the form \begin{equation} (V-I)_{0,g}=a + b \cdot [\mbox{M}/\mbox{H}] + c \cdot [\mbox{M}/\mbox{H}]^2 + d \cdot [\mbox{M}/\mbox{H}]^3 \label{equVIFeHCalib} \end{equation} In addition, we used the $[\mbox{M}/\mbox{H}]$ and $(V-I)_{0,g}$ values for M67 given by Montgomery et al. (\cite{MON93}) to check the zero point. Taking into accound that M67 is even younger than NGC 6791 by 3 to 5 Gyrs, the measured quantities fit reasonably well. Table \ref{tabVIcalib} contains the calibration coefficients, Fig. \ref{diaVIcalib} the graphic relations, again for different ages. As above, we used the relation for $\log(age)=10.160$. \begin{table*}[hb] \footnotesize \begin{center} \begin{tabular}{lcccc} \hline $\log(age)$ & a & b & c & d \\ \hline $9.990$ &$1.279\pm 0.002$ &$0.438 \pm 0.004$ &$0.287\pm 0.010$ &$0.092\pm 0.005$ \\ $10.120$ &$1.315\pm 0.003$ &$0.468 \pm 0.007$ &$0.307\pm 0.016$ &$0.099\pm 0.008$ \\ $10.160$ &$1.330\pm 0.004$ &$0.467 \pm 0.008$ &$0.281\pm 0.021$ &$0.088\pm 0.011$ \\ $10.200$ &$1.343\pm 0.006$ &$0.479 \pm 0.006$ &$0.289\pm 0.014$ &$0.091\pm 0.008$ \\ \hline \end{tabular} \end{center} \normalsize \caption{Calibration coefficients for the $(V-I)_{0,g}-[\mbox{M}/\mbox{H}]$-relation (equation \ref{equVIFeHCalib}).\label{tabVIcalib}} \end{table*} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{8609f31.ps}} \caption{Calibration of the non-linear $(V-I)_{0,g}-[\mbox{M}/\mbox{H}]$-relation for different ages. The key for the symbols is the same as used in Fig. \ref{diaCalibdV}, except for the filled triangle denoting the $(V-I)_{0,g}-[\mbox{M}/\mbox{H}]$ pair of M67. In addition, the quadratic relation of Caretta \& Bragaglia (\cite{CAA98}) is plotted in dash-dotted line. It intersects our relation at low $[Fe/H]$ but shows a difference of $\Delta (V-I)=0.4$ mag in the more metal-rich regime. \label{diaVIcalib}} \end{figure} Using the metallicities listed in Table \ref{tabFeHdV}, column $[\mbox{M}/\mbox{H}]$, we get the absolute reddening as given in Table \ref{tabReddening}. Metallicities as well as reddenings fit very well with the values derived via isochrone-fitting, but are significantly lower than the values given in the literature. This is partly explained by the fact that we take the minimal reddening from the reddening map. Another part of the explanation may be that previous isochrone fits tend to use the red ridge of the RGB and thus overestimate the reddening. \begin{table*}[hb] \begin{center} \begin{tabular}{lcccccc} \hline NGC &$E_{V-I}^{iso}$ &$(V-I)_{HB}$ &$(V-I)_{0,g} $ &$E_{V-I}^{rel}$ &$E_{V-I}^{lit}$ &literature \\\hline 5927 &$0.43\pm 0.02$ &$1.63\pm 0.02$ &$1.22\pm 0.04$ &$0.41 \pm 0.05$ & $0.66$ &Sarajedini \& Norris (\cite{SAN94})\\ 6316 &$0.62\pm 0.03$ &$1.76\pm 0.03$ &$1.13\pm 0.06$ &$0.63 \pm 0.06$ & $0.61$ &Davidge et al. (\cite{DAV92})\\ 6342 &$0.46\pm 0.03$ &$1.65\pm 0.04$ &$1.18\pm 0.05$ &$0.47 \pm 0.06$ & $0.65$ &Armandroff \& Zinn (\cite{ARZ88})\\ 6441 &$0.49\pm 0.03$ &$1.64\pm 0.04$ &$1.19\pm 0.05$ &$0.46 \pm 0.06$ & $0.64$ &Deutsch et al. (\cite{DEU96})\\ 6760 &$0.72\pm 0.03$ &$1.88\pm 0.02$ &$1.14\pm 0.05$ &$0.74 \pm 0.06$ & $1.07$ &Armandroff \& Zinn (\cite{ARZ88})\\ 6528 &$0.46\pm 0.03$ &$1.79\pm 0.04$ &$1.33\pm 0.06$ &$0.46 \pm 0.06$ & $0.70$ &Richtler et al. (\cite{RTL98})\\ 6553 &$0.76\pm 0.03$ &$2.08\pm 0.04$ &$1.31\pm 0.06$ &$0.77 \pm 0.06$ & $0.95$ &Sagar et al. (\cite{SAG98})\\ \hline \end{tabular} \end{center} \caption{Absolute reddening for all GCs. Column $E_{V-I}^{iso}$ contains values derived via isochrone-fitting, column $E_{V-I}^{rel}$ values via $(V-I)_{0,g}-[\mbox{M}/\mbox{H}]$-relation. $(V-I)_{g}$ gives the colour of the RGB at the level of $V_{HB}$, $(V-I)_{0,g}$ the corresponding dereddened colour, calculated via the $(V-I)_{0,g}-[\mbox{M}/\mbox{H}]$-relation. In the last column, we cited values form literature and their sources, which, of course, cannot be more than a selection. Any measurement of colours was done in the differentially dereddened CMDs, which provides the explanation for the difference between our values and that taken from other works. \label{tabReddening}} \end{table*} Sarajedini (\cite{SAR94}) proposed a method to simultaneously determine metallicity and reddening. For this, he used the (linear) $[\mbox{M}/\mbox{H}]-(V-I)_{0,g}$-relation and the dependence of metallicity on the luminosity of the RGB at the absolute colour of $V-I=1.2$ mag, in linear form as well. He calibrated both relations for a metallicity range of $(-2.2 \le [\mbox{M}/\mbox{H}] \le -0.70)$ dex. We recalibrated these relations in order to use them for our clusters. Using NGC 6791 and the Padova-tracks. The graphic results are shown in Figs. \ref{diaSarVI} and \ref{diaSar12}; the calibration coefficients are given in Table \ref{tabSarCalib}. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{8609f32.ps}} \caption{Calibration of the linear $[\mbox{M}/\mbox{H}]-(V-I)_{0,g}$-relation according to Sarajedini (filled symbols) in comparison to the recalibration for higher metallicities. The key for the symbols is the same as used in Fig. \ref{diaCalibdV}, except for the filled symbols. \label{diaSarVI}} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{8609f33.ps}} \caption{Calibration of the linear $[\mbox{M}/\mbox{H}]-\Delta V_{1.2}$-relation according to Sarajedini (filled symbols) in comparison to the recalibration for higher metallicities. The key for the symbols is the same as used in Fig. \ref{diaCalibdV}, except for the filled symbols. \label{diaSar12}} \end{figure} \begin{table*}\ \begin{center} \begin{tabular}{lcccc} \hline & $a_{V-I}$ & $b_{V-I}$ & $a_{1.2}$ & $b_{1.2}$ \\ \hline Padova 7.9 Gyr & $3.432$ & $-4.595$ & $-0.442$ & $-0.340$ \\ Padova 13.2 Gyr & $3.225$ & $-4.451$ & $-0.441$ & $-0.497$ \\ Padova 14.5 Gyr & $3.222$ & $-4.458$ & $-0.407$ & $-0.506$ \\ Padova 15.8 Gyr & $3.052$ & $-4.280$ & $-0.394$ & $-0.543$ \\ Sarajedini & $9.668$ & $10.64$ & $-0.9367$ & $0.2606$ \\ \hline \end{tabular} \end{center} \caption{Coefficients for Sarajedini- and Padova-relations. The equations have the form $[\mbox{M}/\mbox{H}]=a_{V-I}(V-I)_{0,g} + b_{V-I}$ and $[\mbox{M}/\mbox{H}]=a_{1.2}\Delta V_{1.2} + b_{1.2}$. The errors are $\Delta a_{V-I}=0.43$, $\Delta b_{V-I}=0.23$, $\Delta a_{1.2}=0.04$ und $\Delta b_{1.2}=0.16$. \label{tabSarCalib}} \end{table*} For a discussion and new calibration of Sarajedini's method see Caretta \& Bragaglia (\cite{CAA98}). We did not make use of this method, as the extrapolation of Sarajedini's calibration did not seem to be advisable, with reference to Figs. \ref{diaSarVI} and \ref{diaSar12}. \subsubsection{Distance} The brightness $M_V^{HB}$ of the horizontal branch is the best distance indicator for GCs. However, there is a lively discussion on how this brightness depends on the metallicity of the cluster. We take the LMC distance as the fundamental distance for calibrating the zero point in the relation between metallicity and horizontal branch/RR\,Lyrae brightness. The third fundamental distance determination beside trigonometric parallaxes and stellar stream parallaxes is the method of Baade-Wesselink parallaxes. It had been applied to the LMC in its modified form known as Barnes-Evans parallaxes. So far, it has been applied to Cepheids in \object{NGC~1866} (Gieren et~al. \cite{GIE94}), and the most accurate LMC distance until now stems from the period-luminosity relation of LMC Cepheids by Gieren et al. (\cite{GIE98}). We adopt the distance modulus from the latter work, which is $18.46\pm0.06$ mag, and which is in very good agreement with most other work (e.g. Tanvir \cite{TAN96}). If we adopt the apparent magnitude of RR\,Lyrae stars in the LMC from Walker (\cite{WAL92}), $18.94\pm0.1$ mag for a metallicity of $[\element{Fe}/\element{H}]=-1.9$ dex, and the metallicity dependence from Caretta et~al. (\cite{CAB98}), one gets \begin{equation} M_V(RR) = (0.18\pm0.09)([\element{Fe}/\element{H}]+1.6)+0.53\pm0.12 \label{equCaretta} \end{equation} This zero-point is in excellent agreement with the one derived from HB-brightnesses of old LMC globular clusters, if the above metallicity dependence is used (Olszewski et al. \cite{OLS91}). With relation \ref{equCaretta}, with the reddenings (as shown in Table \ref{tabReddening}, column $E_{V-I}^{rel}$) and with the extinction $A_V=R_V^I E_{V-I}$ we can calculate the distance moduli \begin{equation} (m-M)_0=V_{HB}-A_V-M_V^{HB} \label{equDistModHB} \end{equation} The values for $M_V^{HB}$ and $[\mbox{M}/\mbox{H}]$ are listed in Table \ref{tabFeHdV}, and the results are given in Table \ref{tabDistRel}. $R_V^I$ comes from Table \ref{tabDifRed}. \begin{table*}[hb] \begin{center} \begin{tabular}{lccccc} \hline NGC &$A_V$ &$M_V^{HB}$ &$(m-M)_0$ &$r$ &$r_{Harris}$ \\ \hline 5927 &$0.79\pm 0.12$ &$0.77\pm 0.12$ &$14.75\pm 0.17$ &$8.9\pm 0.7$ &$7.4$ \\ 6316 &$1.32\pm 0.18$ &$0.71\pm 0.13$ &$15.39\pm 0.23$ &$12.0\pm 1.2$ &$11.5$ \\ 6342 &$1.04\pm 0.15$ &$0.74\pm 0.13$ &$14.69\pm 0.20$ &$8.7\pm 0.8$ &$9.1$ \\ 6441 &$1.05\pm 0.16$ &$0.75\pm 0.13$ &$15.70\pm 0.21$ &$13.8\pm 1.3$ &$9.7$ \\ 6760 &$1.48\pm 0.13$ &$0.72\pm 0.13$ &$14.88\pm 0.19$ &$9.5\pm 0.8$ &$7.3$ \\ 6528 &$1.11\pm 0.16$ &$0.82\pm 0.12$ &$15.15\pm 0.24$ &$10.7\pm 1.1$ &$7.4$ \\ 6553 &$1.76\pm 0.13$ &$0.81\pm 0.12$ &$13.95\pm 0.24$ &$6.2\pm 0.7$ &$4.7$ \\ \hline \end{tabular} \end{center} \caption{Distances of GCs via $M_V^{HB}-[\mbox{M}/\mbox{H}]$-relation. The given errors only account for the errors in $M_V^{HB}$. The values of column $r_{Harris}$ are taken from Harris (\cite{HAR96}). Distances in kpc, brightness in mag.\label{tabDistRel}} \end{table*} \subsection{Comparison\label{ssecComp}} The distances determined via the $M_{HB}-[\mbox{M}/\mbox{H}]$-relation are larger than those determined by the isochrone-fitting (Table \ref{tabDistComp}). However, as the related reddenings do not show any significant differences, this effect is attributed to the $M_{HB}-[\mbox{M}/\mbox{H}]$-relation and the isochrone-fitting itself. As described above, the isochrone fitting is lacking the desired accuracy especially because the TOP cannot be resolved for most of the clusters. Moreover, the fact that AGB and RGB cannot be distinguished in our CMDs leads to a systematic error in the isochrone distances in the sense that the isochrones tend to have been fitted with brightnesses which are too large. In the following, we discuss some possible explanations for differences between distances taken from the literature and this work. It should be remembered that the distance errors amount to about 10\%. \begin{enumerate} \item The distance to \object{NGC~6528} increases by nearly 30\% compared to Richtler et al. (\cite{RTL98}). Taking into account, that the isochrone (Fig.\ref{dia6528dciso}) might have been fitted slightly too low, we still get a distance of about $8.1$ kpc. Moreover, Richtler et al. determine the absolute reddening via the differentially reddened CMD, which leads to larger values ($0.6 \le E_{V-I} \le 0.8$) compared to our $E_{V-I}=0.46$. Thus the distance modulus decreases by about $0.4$ mag, as equation \ref{equDistModHB} is corrected more strongly on reddening. Finally, the different slopes of the reddening vector have to be regarded. Richtler et al. assume $A_V/E_{V-I}=2.6$, our slope, which we determined via the slope of the HB, amounts to $A_V/E_{V-I}=2.4$. On the whole, we get a difference between Richtler et al. and this work of $0.7$ mag in the distance modulus. \item In the CMDs of \object{NGC~5927} and 6760 the differential reddening becomes noticable especially along the steep part of the RGB, as this is running nearly perpendicularly to the reddening vector. Around the turn over of the AGB/RGB and for its redder part, it leads to an elongation, but not to a broadening of the structures. Fitting an isochrone to the broadened RGB, one generally would use the middle of the RGB as an orientation, as one cannot distinguish between reddening effects and photometric errors in the outer regions. However, the red part of the AGB/RGB approximately keeps its unextinguished brightness. Thus the differential reddening might be overestimated, which leads to decreasing distances. A similar point can be made for the determination of $E_{V-I}$ via the $(V-I)_{0,g}-[\mbox{M}/\mbox{H}]$-relation. Measuring the colour $(V-I)_g$ in the differentially reddened diagram is best done at the middle of the broadened RGB again. This leads to an increased reddening, i.e. the distance modulus will be corrected too strongly for extinction. Overestimating the colour by $0.1$ mag leads to a decrease in distance of about 10\%. \item For \object{NGC~6441}, Harris (\cite{HAR96}) cites a value of $V_{HB}=17.10$ mag. From our CMD we get $17.66$ mag. This lower brightness is supported by (V,B-V)-CMDs of Rich et al. (\cite{RIC97}), especially as Harris' value comes from a CMD by Hesser \& Hartwick (\cite{HES76}), whose lower limiting brightness is around $17.3$ mag. \item The distances as determined via the $[\mbox{M}/\mbox{H}]-\Delta V$- and the $(V-I)_{0,g}-[\mbox{M}/\mbox{H}]$-relation relate to the differentially dereddened CMDs, i.e to the minimal absolute reddening. However, the papers we obtained the cited values from (Table \ref{tabReddening}), do not take differential reddening into account (e.g. Armandroff (\cite{ARM88}) for \object{NGC~6342} and 6760, Ortolani et al. (\cite{ORT90}, \cite{ORT92}) for \object{NGC~6528} and 6553). So their absolute reddenings are systematically larger and the distances smaller. Interestingly, the absolute reddening for \object{NGC~6316} of $E_{V-I}=0.63$ mag as determined in this work fits very well the value of $E_{V-I}=0.61$ mag given by Davidge et al. (\cite{DAV92}); \object{NGC~6316} shows the smallest differential reddening ($\delta E_{V-I}=0.07$ mag) of our cluster sample. \item Finally, the distances depend on the assumed extinction law. The value varies between $3.1 \le R_V^B \le 3.6$ (Savage \& Mathis \cite{SAV79}, Grebel \& Roberts \cite{GRE95}, see Fig \ref{diaExtVar} and discussion). This effect should have the strongest influence on the distances as determined in this work. Taking an absolute reddening of $0.5$ mag, the variation between the above cited values results in a difference of $0.25$ mag in the distance modulus. This corresponds to about 25\% of the distance in kpc. \end{enumerate} The distance error mostly depends on the absolute reddening used. The errors in the metallicites have only a minor influence on the distances (see Table \ref{tabVIcalib}). They amount to around 3\% of the total distance in kpc. In conclusion, the increased distances $r_{rel}$ (Table \ref{tabDistComp}) are due to the fact that we determine the distance-relevant parameters using the differentially dereddened CMDs. \begin{table}[hb] \begin{center} \begin{tabular}{lccccc} \hline NGC &$r_{Harris}$ &$r_{iso}$ &$r_{rel}$ &$r_{lit}$ \\ \hline 5927 &$ 7.4$ &$ 8.4$ &$ 8.9$ & \\ 6316 &$11.5$ &$12.3$ &$12.0$ & \\ 6342 &$ 9.1$ &$ 7.4$ &$ 8.7$ & \\ 6441 &$ 9.7$ &$11.8$ &$13.9$ & \\ 6760 &$ 7.3$ &$ 8.9$ &$ 9.5$ & \\ 6528 &$ 7.4$ &$ 9.3$ &$10.7$ &$6.6$ \\ 6553 &$ 4.7$ &$ 5.4$ &$ 6.2$ &$5.2$ \\ \hline \end{tabular} \end{center} \caption{Comparison of the distances taken from Harris (\cite{HAR96}), $r_{Harris}$, with the values of this work. $r_{iso}$ and $r_{rel}$ contain the distances determined via isochrone-fitting and $M_{HB}-[\mbox{M}/\mbox{H}]$-relation. The last column shows recently determined distances for \object{NGC~6528} (Richtler et al. \cite{RTL98}) and for \object{NGC~6553} (Guarnieri et al. \cite{GUA98}). \label{tabDistComp}} \end{table} \clearpage \subsection{Masses} To classify the clusters according to Burkert \& Smith (\cite{BUR97}), we have to determine the masses from the total absolute brightnesses. Because we could not measure the apparent total brightness, we used the values given by Harris (\cite{HAR96}). With the extinctions and distance moduli given above (Table \ref{tabDistRel}), we get the absolute total brightnesses via \begin{equation} M_V^{total}=V^{total}- A_V - (m-M)_0 \end{equation} We determined the masses using a mass-to-light-ratio of $\left(\frac{M}{L}\right)_V=3$ (Chernoff \& Djorgovski \cite{CHE89}). Table \ref{tabMasses} shows the results. Thus, \object{NGC~6441} is one of the most massive clusters of the galaxy. $\omega$ Cen/NGC 5139 has $\log(M/M_{\odot})= 6.51$ (Harris \cite{HAR96}). \begin{table}[hb] \begin{center} \begin{tabular}{lccc} \hline NGC & $V^{total}$ & $M_V^{total}$ & $\log\left(\frac{M}{M_{\odot}}\right)$ \\ \hline 5927 & $8.01$ & $-7.52$ & $5.42$ \\ 6316 & $8.43$ & $-8.28$ & $5.72$ \\ 6342 & $9.66$ & $-6.07$ & $4.85$ \\ 6441 & $7.15$ & $-9.61$ & $6.26$ \\ 6760 & $8.88$ & $-7.48$ & $5.41$ \\ 6528 & $9.60$ & $-6.65$ & $5.07$ \\ 6553 & $8.06$ & $-7.65$ & $5.47$ \\ \hline \end{tabular} \end{center} \caption{Absolute total brightnesses and masses for all clusters. The apparent brightnesses were taken from Harris (\cite{HAR96}). \label{tabMasses}} \end{table} \section{Classification and assignment} After having determined the parameters of our cluster sample, we now discuss each cluster's possible affiliations with the galactic structure components i.e. halo, disk or bulge for each cluster. The necessary criteria are introduced in the following subsection. \subsection{The assignment criteria } \subsubsection{Disk and Halo: Zinn (1985)} Zinn (\cite{ZIN85}) divided the GC-system into a metal-poor ($[\mbox{M}/\mbox{H}] \le -0.8$ dex) halo- and a metal-rich ($[\mbox{M}/\mbox{H}] \ge -0.8$) disk-subsystem. This distinction also correlated with the kinematics and spatial distribution of their objects. The resulting criteria are listed in Table \ref{tabZinn}. Equation \ref{equVrad} gives the orbital velocity $v_c$ of a cluster depending on its observed radial velocity $v_{rad}$. $v_c$ can be compared to the net rotation as given in Table \ref{tabZinn}. \begin{table}[hb] \begin{center} \begin{tabular}{l|ccc} \hline subsystem &$[\mbox{M}/\mbox{H}]\le -0.8$dex &$[\mbox{M}/\mbox{H}]\ge -0.8$dex \\ \hline $v_{rot}$[km/s] &$50 \pm 23$ &$152 \pm 29$ \\ $\sigma_{rot}$[km/s]&$114$ &$71$ \\ \hline \end{tabular} \end{center} \caption{Kinematics and spatial distribution of the metal-rich and -poor subsystems of GCs according to Zinn (1985) \label{tabZinn}} \end{table} \subsubsection{Bulge and (thick) disk: Minniti (1995,1996)} Minniti (\cite{MIN95}, \cite{MIN96}) divided Zinn's metal-rich disk system further into GCs belonging to the (thick) disk on the one hand and to the bulge on the other. Comparing the GCs with their corresponding field population, he assigned the GCs with galactocentric distances $R_{gc} \le 3$ kpc to the bulge and the ones with $R_{gc} \ge 3$ kpc to the thick disk. \subsubsection{Inner halo, bar and disk: Burkert \& Smith} Burkert \& Smith (\cite{BUR97}) used the masses of the metal-rich GCs to distinguish between a group belonging to the inner halo and a group which can be further divided into a bar- and a ring-system using the kinematics and spatial distribution of the clusters (see Table \ref{tabBurkert}). \begin{table}[hb] \begin{center} \begin{tabular}{l|ccc} \hline group & inner halo & bar & disk \\ \hline $\log(M/M_s)$ & $\ge 5.55$ & $\le 5.55$ & $\le 5.55$ \\ $v_{rot}$[km/s] & $24\pm 23$ & $24\pm 23$ & $164\pm 6$ \\ $v_{rot}/\sigma_{rot}$ & $0.3$ & $0.3$ & $6$ \\ spatial & concentrated & bar-like & $(4\le R_{gc}\le 6)$kpc \\ distribution & to center & structure & \\ \hline \end{tabular} \caption{Criteria for subgroups of the metal-rich GCs according to Burkert \& Smith (\cite{BUR97}). \label{tabBurkert}} \end{center} \end{table} \subsubsection{Radial velocities} Unfortunately, there do not exist any data on proper motions of our clusters. The only kinematic information available are radial velocities, catalogued by Harris (\cite{HAR96}). Thus, we can only check, whether a disk orbit is compatible with a given radial velocity. This is possible by comparing the measured radial velocity $v_{rad}$ with the expected one, calculated via equation \ref{equVrad} assuming that disk clusters move on circular orbits in the galactic plane. \begin{equation} v_{rad}=v_c\sin \left(l+\arctan\left(\frac{y}{R_s-x}\right)\right) -v_s\sin (l), \label{equVrad} \end{equation} where $l$ is the galactic longitude, and $x$ and $y$ are the heliocentric coordinates. We used $R_s=8.0$ kpc and $v_s=220$ km/s. $v_c$ gives the velocities of the clusters in the plane, corresponding to the galactic rotational velocity $v_{rot}(R_s-x)$ with the values taken from Fich \& Tremaine (\cite{FIC91}). \subsubsection{Metallicity gradient} The metallicity gradient of the disk is an uncertain criteria insofar, as it is defined for the outer ranges of the galactic disk. We use a metallicity gradient referring to the population of old open clusters. The oldest of these objects have ages similar to the youngest GCs (Phelps et al. \cite{PHE94}). Their scale height is comparable to other thick disk objects. Assuming that they are related to a possible disk population of GCs (Friel \cite{FRI95}), we can use their metallicity gradient \begin{equation} \frac{\partial [\mbox{Fe}/\mbox{H}]}{\partial R_{gc}}=-0.091 \pm 0.014 \end{equation} (Friel \cite{FRI95}) as a criterion for whether our GCs belong to the galactic thick disk or not. \subsection{The assignment} Using the above criteria, we assigned the clusters of our sample according to Table \ref{tabAssign}. The values of the parameters necessary to decide on group membership are listed in Table \ref{tabAssDat}. \begin{table}[hb] \footnotesize \begin{center} \begin{tabular}{l|ccccc} NGC & Zinn & Minniti & Burkert & $v_{rad}^{dsk}$ & $\partial_r[\mbox{M}/\mbox{H}]$ \\ \hline 5927 & d & (d) & d & +d & -d \\ 6316 & ? & ? & bu & +d & -d \\ 6342 & ? & bu & ba & -d & -d \\ 6441 & ? & ? & bu & -d & -d \\ 6760 & d & ? & d & -d & -d \\ 6528 & ? & bu & ba & -d & -d \\ 6553 & (h) & bu & ba & -d & -d \\ \hline \end{tabular} \end{center} \normalsize \caption{Assignment of the clusters to the systems \textbf{d}isk, \textbf{h}alo, \textbf{bu}lge and \textbf{ba}r according to the criteria in the first column. \textbf{?} is used if no assignment is possible, the symbols d, h, bu and ba together with \textbf{+} or \textbf{-} relate to criteria which only can decide whether an object belongs to a certain group or not. Symbols in brackets denote uncertainties explained in the text. \label{tabAssign}} \end{table} \begin{table*} \begin{center} \begin{tabular}{l|c|c|c|c|c|c|c} \hline NGC &5927 &6316 &6342 &6441 &6760 &6528 &6553 \\ \hline $[\mbox{M}/\mbox{H}]$[dex] &$-0.29$ &$-0.60$ &$-0.42$ &$-0.39$ &$-0.57$ &$0.00$ &$-0.04$ \\ $R_{gc}$[kpc] &$5.0$ &$4.1$ &$1.7$ &$6.1$ &$5.6$ &$2.8$ &$2.0$ \\ $z$[kpc] &$0.8$ &$1.3$ &$1.6$ &$-1.3$ &$-0.7$ &$-0.8$ &$-0.3$ \\ $\log(M/M_{\odot})$ &$5.42$ &$5.72$ &$4.84$ &$6.26$ &$5.41$ &$5.07$ &$5.47$ \\ $v_{rot}$[km/s] &$220$ &$220$ &$210$ &$225$ &$220$ &$195$ &$210$ \\ $v_{rad}$[km/s] &$-116$ &$72$ &$81$ &$18$ &$-28$ &$185$ &$-7$ \\ $v_{rot}^{dsk}$[km/s] &$265$ &$-308$ &$114$ &$18$ &$121$ &$1910$ &$37$ \\ $v_{rad}^{dsk}$[km/s] &$-75$ &$-32$ &$164$ &$-58$ &$56$ &$15$ &$58$ \\ $[\mbox{M}/\mbox{H}]_{oc}(r)$[dex]&$0.18$ &$0.20$ &$0.46$ &$0.01$ &$0.11$ &$0.33$ &$0.51$ \\ \hline \end{tabular} \end{center} \caption{All relevant parameters for the assignment. $[\mbox{M}/\mbox{H}]$ gives the metallicity according to the $[\mbox{M}/\mbox{H}]-\Delta V$-relation, $R_S$ the heliocentric distance, $R_{gc}$ the galactocentric distance and $z$ the distance to the disk. The M in $\log(M/M_{\odot})$ stands for cluster mass, $v_{rot}$ for the orbital velocity, derived via the rotational velocity curve of Fich \& Tremaine (\cite{FIC91}), $v_{rad}$ for the observed radial velocity and $v_{rot}^{dsk}$ for the orbital velocity of the clusters, calculated using $v_{rad}$ and the assumption of circular cluster orbits (in order to compare with a net rotation). $v_{rad}^{dsk}$ contains the expected radial velocity assuming disk orbits and $[\mbox{M}/\mbox{H}]_{oc}(r)$ the values derived via a metallicity gradient of the old open clusters (Friel \cite{FRI95}). \label{tabAssDat}} \end{table*} As for the metallicities of our clusters, they all belong to the disk system according to Zinn, which is obvious as the sample had been selected in this way. Not so obvious is the comparison with the net rotation of Zinn's disk group. Only \object{NGC~5927} shows a value of $v_{rot}$ which is not totally off the net rotation as given in Table \ref{tabZinn}. The clusters belonging to the bulge according to Minniti's criterion are members of the bar following the arguments of Burkert \& Smith (\cite{BUR97}). Binney et al. (\cite{BIN97}) quote a value of $20^{\circ}$ for the angle between x-axis in galactocentric coordinates and the major semiaxis of the bulge structure. Its end lying nearer to the sun is located at small galactic longitudes ($y \le 0$ in cartesian coordinates). Fig \ref{diaXYZ} shows the spatial distribution of our cluster sample. The coordinates of the 'bar' clusters \object{NGC~6342}, 6528 and 6553 according to Burkert \& Smith seem to be consistent with a structure described by Binney et al. (\cite{BIN97}). However, as the referee pointed out, we do not know how long-lived the Milky-Way bar is, and other tracers of old populations such as RR Lyrae do not follow the bar (Alcock et al. \cite{ALC98}). Moreover, the distance between the 'bar' clusters \object{NGC 6528} and \object{NGC 6553} is about $5$ kpc, which is much larger than the length of the Milky-Way bar according to most authors (e.g. Binney et al. \cite{BIN97}). Also note in Fig \ref{diaXYZ}, that the errors in the x-coordinate are larger than those in y and z. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{8609f34.ps}} \caption{Distribution of the cluster sample in the galactocentric x-y-plane (a) and y-z-plane (b). In the upper panel, the observer is located at $z \ge 0$. The distribution in the y-z-plane is as seen from the sun. The distances used correspond to $r_{rel}$ in Table \ref{tabDistComp}.\label{diaXYZ}} \end{figure} There are only two 'disk' clusters remaining, assuming Burkert \& Smith's definition of disk clusters: \object{NGC~5927} and \object{NGC~6760}. However, the radial velocities corroborate this result for \object{NGC~5927} only. For any other cluster, the radial velocities seem to exclude an assignment to the disk. The metallicity gradient of the old open clusters leads to the conclusion that none of our clusters is to be assigned to the thick disk. Taken the whole sample of metal rich clusters (i.e. clusters with $[\mbox{M}/\mbox{H}] \ge -0.8$ dex according to Zinn \cite{ZIN85}, see Tables \ref{tabRestAss}, \ref{tabRestDat}), we find only three objects, which could be disk clusters according to the metallicity gradient criterium. \begin{table*} \footnotesize \begin{center} \begin{tabular}{lcccccccc} \hline Name &$R_{gc}$&$z$ &$[\mbox{M}/\mbox{H}]$&$v_{rad}$&$v_{rad}^{dsk}$&$\log(M/M_{\odot})$&$v_{rot}^{dsk}$&$[\mbox{M}/\mbox{H}]_{oc}$\\ \hline NGC 104 &$7.3$ &$-3.0$&$-0.76$ &$ -19$ &$ -26$ &$6.16$ &$202$ &$0.06$ \\ Lynga 7 &$4.2$ &$-0.3$&$-0.62$ &$ 8$ &$ -95$ & &$106$ &$0.35$ \\ NGC 6256 &$2.3$ &$ 0.5$&$-0.70$ & & &$4.87$ & &$0.52$ \\ NGC 6304 &$2.2$ &$ 0.6$&$-0.59$ &$-107$ &$ -38$ &$5.32$ &$478$ &$0.52$ \\ NGC 6356 &$7.0$ &$ 2.6$&$-0.50$ &$ 27$ &$ -56$ &$5.81$ &$-361$ &$0.09$ \\ NGC 6352 &$3.3$ &$-0.7$&$-0.70$ &$-121$ &$ -93$ &$4.99$ &$244$ &$0.42$ \\ Ter 2 &$1.6$ &$ 0.4$&$-0.25$ &$ 109$ &$ 79$ &$4.41$ &$305$ &$0.58$ \\ Liller 1 &$2.6$ &$ 0.0$&$ 0.22$ &$ 53$ &$ 73$ &$5.45$ &$127$ &$0.49$ \\ Ter 1 &$1.5$ &$ 0.1$&$-0.35$ &$ 35$ &$ -40$ &$3.71$ &$-108$ &$0.59$ \\ Ton 2 &$1.4$ &$-0.5$&$-0.50$ & & &$4.84$ & &$0.60$ \\ NGC 6388 &$4.4$ &$-1.4$&$-0.60$ &$ 81$ &$ 155$ &$6.32$ &$ 54$ &$0.32$ \\ Pal 6 &$1.4$ &$ 0.2$&$-0.10$ &$ 201$ &$ 31$ &$5.34$ &$1112$ &$0.60$ \\ Ter 5 &$0.6$ &$ 0.2$&$-0.28$ &$ -94$ & &$5.56$ & &$0.67$ \\ NGC 6440 &$1.2$ &$ 0.5$&$-0.34$ &$ -79$ &$ 180$ &$5.89$ &$-49$ &$0.62$ \\ Ter 6 &$0.6$ &$-0.3$&$-0.65$ &$ 126$ &$ -77$ &$5.14$ &$-305$ &$0.67$ \\ NGC 6496 &$4.4$ &$-2.0$&$-0.64$ &$ 113$ &$ 136$ &$5.28$ &$-364$ &$0.32$ \\ Ter 10 &$0.8$ &$-0.3$&$-0.70$ & & &$5.52$ & &$0.66$ \\ NGC 6539 &$3.0$ &$ 0.9$&$-0.66$ &$ -45$ &$ 130$ &$5.71$ &$ 32$ &$0.46$ \\ NGC 6624 &$1.2$ &$-1.1$&$-0.42$ &$ 53$ &$ 181$ &$5.39$ &$ 71$ &$0.62$ \\ NGC 6637 &$1.5$ &$-1.5$&$-0.71$ &$ 40$ &$-191$ &$5.40$ &$-52$ &$0.59$ \\ Pal 8 &$5.2$ &$-1.5$&$-0.48$ &$ -43$ &$-134$ &$4.60$ &$-27$ &$0.26$ \\ Pal 10 &$6.4$ &$ 0.3$&$-0.10$ & & &$4.71$ & &$0.14$ \\ Pal 11 &$7.6$ &$-3.4$&$-0.39$ &$ -68$ &$ -24$ &$5.11$ &$-72$ &$0.04$ \\ NGC 6838 &$6.7$ &$-0.3$&$-0.73$ &$ -23$ &$ 25$ &$4.62$ &$161$ &$0.12$ \\ \hline \end{tabular} \normalsize \end{center} \caption{The relevant parameters of the remaining metal rich GCs according to Harris (\cite{HAR96}). The columns are labeled as in Table \ref{tabAssDat}. Distances are given in kpc, metallicities in dex and velocities in km/s. \label{tabRestDat}} \end{table*} \begin{table} \footnotesize \begin{center} \begin{tabular}{lcccccc} \hline Name &Zinn & Minniti & Burkert &$v_{rad}$ &$[\mbox{M}/\mbox{H}]_{oc}$ \\\hline NGC 104 &d &- &? &+d &-d \\ Lynga 7 &d &- &- &-d &-d \\ NGC 6256 &- &bu &ba &-d &-d \\ NGC 6304 &? &bu &ba &-d &-d \\ NGC 6356 &? &- &? &-d &-d \\ NGC 6352 &? &- &ba &+d &-d \\ Ter 2 &? &bu &ba &+d &-d \\ Liller 1 &d &bu &ba(d) &+d &+d \\ Ter 1 &? &bu &ba &-d &-d \\ Ton 2 &? &bu &ba &- &-d \\ NGC 6388 &d &- &? &d? &-d \\ Pal 6 &?(bu)&bu &ba &-d &-d \\ Ter 5 &- &bu &bu &- &-d \\ NGC 6440 &? &bu &bu &-d &-d \\ Ter 6 &? &bu &bu &-d &-d \\ NGC 6496 &? &- &ba &+d &-d \\ Ter 10 &? &bu &ba &- &-d \\ NGC 6539 &? &bu? &bu &-d &-d \\ NGC 6624 &? &bu &ba &-d &-d \\ NGC 6637 &? &bu &ba &-d &-d \\ Pal 8 &? &- &d &-d &-d \\ Pal 10 &- &-(d) &d &- &+d \\ Pal 11 &? &? &d &d? &+d? \\ NGC 6838 &d &- &d &-d &-d \\ \hline \end{tabular} \normalsize \end{center} \caption{Suggested assignment according to the criteria discussed above for the remaining metal rich GCs. The columns are labeled as in Table \ref{tabAssign}. \label{tabRestAss}} \end{table} Some of the clusters do not meet any of the criteria. Interestingly, they are the most massive, but metal-poorest objects of the sample. These objects are \object{NGC~6316}, 6760, and 6441 as well as (in Table \ref{tabRestAss}) NGC 104, 6356 and 6388. Although NGC 104 seems to be a disk cluster and mostly is referred to as such, the large distance to the galactic plane (3 kpc) does not support this assignment. Probably, the mentioned objects belong to the halo, being its metal-richest clusters. Zinn (\cite{ZIN85}) and Armandroff (\cite{ARM93}) point to the fact that the division into metal-rich and poor clusters is by no means an exact one, but that there is a metal-rich sample of halo clusters as well as a metal-poorer one of disk objects. Richtler et al. (\cite{RTL94}) discussed the existence of a subgroup of disk clusters according to Zinn (\cite{ZIN85}), based on an analysis of the metallicities and the minimum inclination angles derived from $z/R_{gc}$-values for these clusters. They conclude that the clusters NGC 6496, 6624 and 6637 might not be disk clusters after all, but belong to the halo. Adding their argument to the above discussion, we end up with 3 probable disk members (\object{NGC~5927}, additionally from Table \ref{tabRestAss} Liller 1 and Pal 10. Pal 11 is excluded because of its large minimum inclination angle.) and 9 clusters (NGC 104, 6316, 6356, 6388, 6441, 6496, 6624, 6637 and 6760) that more likely belong to the halo than to the (thick) disk. The rest of the clusters (\object{NGC~6342}, 6528, 6553 and the remaining ones of Table \ref{tabRestAss}) fall in with the bulge/bar-group of Minniti (\cite{MIN95}) and Burkert \& Smith (\cite{BUR97}). \section{Conclusions} We derived the parameters for five GCs near the galactic center in a uniform manner, employing a new calibration of methods which relate structures in the CMDs with the parameters. Taking into account the differential reddening and correcting the CMDs for it leads to more accurately determined parameters and a decreasing absolute reddening. There might be a systematic effect on distances, if the differential reddening is not taken care of. With the $[\mbox{M}/\mbox{H}]-\Delta V$- method we present a accurate way to estimate differentially metallicities of metal-rich GCs. Especially it might prove useful for surveys of clusters in V, V-I, as their CMDs only need to contain the HB and turnover of the AGB/RGB. The metallicities of our program clusters all lie in the range of the clusters constituting the classical disk-system of GCs in the Milky Way. However, different criteria defining subgroups of the GC-system partly lead to differing results. Most of the metal-rich GCs seem to belong to a bar/bulge-structure, and only a minority could clearly be addressed as 'disk'-clusters. So the classical disk-system is more likely to be a mixture between a halo- and bulge-component. \begin{acknowledgements} We are indebted to E.K. Grebel for the most interesting and valuable discussions, especially on the technique of differential dereddening. We would like to thank the referee D. Minniti for helpful comments and criticism. \end{acknowledgements}
\section{ Introduction} It is a great honor to have been awarded the Dannie Heineman Prize in Mathematical Physics and it is a great pleasure to be be in the company today of my fellow prize winners, C.N. Yang, this year's winner of the Lars Onsager prize, A.B. Zamolodchikov, the inventor of conformal field theory, and especially my collaborator and thesis adviser T.T. Wu. \section{Why integrable models is the invisible field of physics} The citation of this year's Heineman Prize is for ``groundbreaking and penetrating work on classical statistical mechanics, integrable models and conformal field theory'' and in this talk I plan to discuss the work in statistical mechanics and integrable models for which the award was given. But since the Heineman prize is explicitly for publications in mathematical physics I want to begin by examining what may be meant by the phrase ``mathematical physics.'' The first important aspect of the term ``mathematical physics'' is that it means something very different from most other kinds of physics. This is seen very vividly by looking at the list of the divisions of the APS. Here you will find 1) astrophysics, 2) atomic physics, 3) condensed matter physics, 4) nuclear physics, and 5) particles and fields but you will not find any division for mathematical physics. This lack of existence of mathematical physics as a field is also reflected in the index of Physical Review Letters where mathematical physics is nowhere to be found. Therefore we see that the Heineman prize in mathematical physics is an extremely curious award because it honors achievements in a field of physics which the APS does not recognize as a field of physics. This lack of existence of mathematical physics as a division in the APS reflects, in my opinion, a deep uneasiness about the relation of physics to mathematics. An uneasiness I have heard echoed hundreds of times in my career in the phrase ``It is very nice work but it is not physics. It is mathematics''. A phrase which is usually used before the phrase ``therefore we cannot hire your candidate in the physics department.'' So the first lesson to be learned is that mathematical physics is an invisible field. If you want to survive in a physics department you must call yourself something else. So what can we call the winners of the Heineman prize in mathematical physics if we cannot call them mathematical physicists? The first winner was Murray Gell--Mann in 1959. He is surely belongs in particles and fields; the 1960 winner Aage Bohr is surely a nuclear physicist; the 1976 winner Stephen Hawking is an astrophysicist. And in fact almost all winners of the prize in mathematical physics can be classed in one of the divisions of the APS without much confusion. But there are at least two past winners who do not neatly fit into the established categories, Elliott Lieb, the 1978 recipient, and Rodney Baxter, the 1987 recipient, both of whom have made outstanding contributions to the study of integrable models in classical statistical mechanics---the same exact area for which Wu and I are being honored here today. This field of integrable models in statistical physics is the one field of mathematical physics which does not fit into some one of the existing divisions of the APS. It is for this reason that I have described integrable models as a hidden field. Indeed it is so hidden that it is sometimes not even considered it to be statistical mechanics as defined by the IUPAP. The obscurity of the field of integrable statistical mechanics models explains why there are less than a dozen physics departments in the United States where it is done. This makes the job prospects of a physicist working in this field very slim. But on the other hand it means that we in the field get to keep all the good problems to ourselves. So it is with mixed feelings that I will now proceed to discuss some of the progress made in the last 33 years and the some of the directions for future research. \section{The Ising Model} The award of the 1999 Heineman prize, even though the citation says that it is for work on integrable models, is in fact for work done from 1966--1981 on a very specific system: the two dimensional Ising model. This work includes boundary critical phenomena, randomly layered systems, the Painlev{\' e} representation of the two point function, and the first explicit results on the Ising model in a magnetic field. All of these pieces of work had results which were unexpected at the time and all have lead to significant extensions of our knowledge of both statistical mechanics and mathematics. \subsection{The boundary Ising model and wetting } The Ising model is a two dimensional collection of classical ``spins'' $\sigma_{j,k}$ which take on the two values $+1$ and $-1$ and are located at the $j$ row and $k$ column of a square lattice. For a translationally invariant system the interaction energy of this system is \begin{equation} {\cal E}= -\sum_{j,k}(E^v\sigma_{j,k}\sigma_{j+1,k}+E^h\sigma_{j,k}\sigma_{j,k+1} +H\sigma_{j,k}). \label{ising} \end{equation} This is certainly one of the two most famous and important models in statistical mechanics (the Heisenberg-Ising chain being the other) and has been studied by some of the most distinguished physicists of this century including Onsager \cite{ons}who computed the free energy for $H=0$ in 1944 and Yang \cite{yanga} who computed the spontaneous magnetization in 1952. In 1966 I began my long involvement with this model when, for part of my thesis, Prof. Wu suggested that I compute for an Ising model on a half plane the same quantities which Onsager and Yang computed for the bulk. At the time we both thought that because the presence of the boundary breaks translational invariance the boundary computations would be at least as difficult as the bulk computations. It was therefore quite surprising when it turned out that the computations were drastically simpler \cite{mccoya}. In the first place the model could be solved in the presence of a field on the boundary which meant that the computation of the magnetization came along for free once we could do the free energy but more importantly the correlation functions, which in the bulk were given by large determinants whose size increased with the separation of the spins, were here given by nothing worse than the product of two one--dimensional integrals for all separations. The key to this great simplification is the fact that the extra complication of the boundary magnetic field actually makes the problem simpler to solve (a realization I had in a dream at 3:00 AM after a New Years eve party). This model is the first case where boundary critical exponents were explicitly computed. Indeed it remained almost the only solved problem of a boundary critical phenomena until it was generalized to integrable massive boundary field theory in 1993 \cite{gz}. This boundary field had the added virtue that we could analytically continue the boundary magnetization into the metastable region and explicitly compute a boundary hysteresis effect \cite{mccoyb}. This leads to the lovely effect that near the value of the boundary magnetic field where the hysteresis curve ends that the spins a very long distance from the boundary turn over from pointing in the direction of the bulk magnetization to pointing in the direction of the metastable surface spin. At the value where the metastability ends this surface effect penetrates all the way to infinity and ``flips'' the spin in the bulk. In later years this phenomena has been interpreted at a ``wetting transition'' \cite{fisher} and the Ising intuition has been extended to many models where exact solutions do not exist. \subsection{Random layered Ising model and Griffiths-McCoy singularities} Our next major project was to generalize the translationally invariant interaction (\ref{ising}) to a non translationally invariant problem where not just a half plane boundary was present but to a case where \begin{enumerate} \item The interaction energies $E^v$ were allowed to vary from row to row but translational invariance in the horizontal direction was preserved \item The interaction energies $E^v(j)$ between the rows were chosen as independent random variables with a probability distribution $P(E^v).$ \end{enumerate} This was the first time that such a random impurity problem had ever been studied for a system with a phase transition and the entire computation was a new invention \cite{mccoyc}. In particular we made the first use in physics of Furstenburg's theory \cite{furs} of strong limit theorems for matrices. We felt that the computation was a startling success because we found that for any probability distribution, no matter how small the variance, there was a temperature scale, depending on the variance, where there was new physics that is not present in the translationally invariant model. For example the divergence in the specific heat at $T_c$ decreases from the logarithm of Onsager to an infinitely differentiable essential singularity. Moreover the average over the distribution $P(E^v)$ of the correlation functions of the boundary could be computed \cite{mccoyd} and it was seen that there was an entire temperature range surrounding $T_c$ where the boundary susceptibility was infinite. Thus the entire picture of critical exponents which had been invented several years before to describe critical phenomena in pure systems was not sufficient to describe these random systems \cite{mccoye}. We were very excited. But then something happened which I found very strange. Instead of attempting to further explore the physics we had found, arguments were given as to why our effect could not possibly be relevant to real systems. This has lead to arguments which continue to this day. We were, and are, of the opinion that the effects seen in the layered Ising model are caused by the fact that the zeroes of the partition function do not pinch the real temperature axis at just one point but rather pinch in an entire line segment. A closely related effect was simultaneously discovered by Griffiths \cite{griffiths} and it was, and is, our contention that in this line segment there is a new phase of the system. But this line segment is not revealed by approximate computations and for decades it was claimed that our new phase was limited to layered systems; a claim which some continue to make to this day. However, fortunately for us, there is an alternative interpretation of the Ising model in terms of a one--dimensional quantum spin chain in which our layered classical two--dimensional system becomes a randomly impure quantum chain \cite{sm}. In this interpretation there is no way to argue away the existence of our new phase and finally in 1995, a quarter century after we first found the effect, D. Fisher \cite{dfisher}, in an astounding paper, was able to craft a theory of the physics of rare events based on an exact renormalization group computation which not only reproduced our the results of our layered model on the boundary but extended the computations to bulk quantities which in 1969 we had been unable to compute. With this computation I think the existence of what are now called Griffiths-McCoy singularities is accepted, but it has taken a quarter century for this to happen. \subsection{Painlev{\' e} Functions and difference equations} But perhaps the most dramatic discoveries were published from 1973 to 1981 on the spin correlation functions of the Ising model and most particularly the results that in the scaling limit where $T\rightarrow T_c$ and the separation of the spins $N \rightarrow \infty$ such that $|T-T_c|N=r$ is fixed that the correlation function \cite{mccoyf} $<\sigma_{0,0}\sigma_{0,N}>$ divided by $|T-T_c|^{1/4}$ is \begin{equation} G_{\pm}(2r)=(1\mp\eta)\eta^{-1/2}{\rm exp}\int_r^{\infty}dx{1\over 4}x\eta^{-2}[(1-\eta^2)^2-\eta'^2] \label{scaling} \end{equation} where $\eta(x)$ satisfies the third equation of Painlev{\' e} \cite{pain} \begin{equation} {d^2\eta\over dx^2}={1\over \eta}({d\eta\over dx})^2-{1\over x}{d\eta\over dx}+{1\over x}(\alpha\eta^2+\beta)+\gamma\eta^3+{\delta\over \eta} \label{pain} \end{equation} with $\alpha=\beta=0$ and $\gamma=-\delta=1$ and \begin{equation} \eta(x)\sim 1-{2\over \pi}K_0(2x)~~~{\rm as}~~x\rightarrow \infty \end{equation} where $K_0(2x)$ is the modified Bessel function of order zero. Furthermore on the lattice all the correlation functions satisfy quadratic difference equations \cite{mccoyg}. This discovery of Painlev{\' e} equations in the Ising model was the beginning of a host of developments in mathematical physics which continues in an ever expanding volume to this day. It led Sato, Miwa, and Jimbo \cite{smj} to their celebrated series of work on isomonodromic deformation and to the solution of the distribution of eigenvalue of the GUE random matrix problem \cite{jmms} in terms of a Painlev{\'e} V function. This has subsequently been extended by many people, including one of our original collaborators, Craig Tracy \cite{tw}, to so many branches of physics and mathematics including random matrix theory matrix models in quantum gravity and random permutations that entire semester long workshops are now devoted to the subject. Indeed a recent book on special functions \cite{iksy} characterized Painlev{\' e} functions as ''the special functions of the 21st century.'' Rarely has the solution to one small problem in physics had so many ramifications in so many different fields. \subsection{ Ising model in a field} The final piece of work in the Ising model to be mentioned is what happens to the two point function when a small magnetic field is put the system for $T<T_c.$ At $H=0$ for $T<T_c$ the two point function has the important property that it couples only to states with an even number of particles and thus, in particular the leading singularity in the Fourier transform is not a single particle pole but rather a two particle cut. In 1978 \cite{mccoyh}, as an application of our explicit formulas for the $n$ spin correlation functions \cite{nspin} we did a (singular) perturbation computation to see what happens when in terms of the scaled variable $h=H/|T-T_c|^{15/8}$ a small value of $h$ is applied to the system. We found that the two particle cut breaks up into an infinite number of poles which are given by the zeroes of Airy functions. These poles are exactly at the positions of the bound states of a linear potential and are immediately interpretable as a weak confinement of the particles which are free at $H=0.$ This is perhaps the earliest explicit computation where confinement is seen. From this result it was natural to conjecture that as we take the Ising model from $T>T_c,~H=0$ to $T<T_c,~H=0$ that as $h$ increases from $0$ to $\infty~(T=T_c,H>0)$ that bound states emerge from the two particle cut and that as we further proceed from $T=T_c,~H>0$ down to $T<T_c,~H=0$ that bound states continue to emerge until at $H=0$ an infinite number of bound states have emerged and formed a two particle cut. What this picture does not indicate is the remarkable result found 10 years later by A. Zamolodchikov \cite{zamb} that at $T=T_c,~ H=0$ the problem can again be studied exactly. This totally unexpected result will be discussed by Zamolodchikov in the next presentation. \section{From Ising to integrable} Even at the time when this Ising model work was initiated there were other models known such as the Heisenberg chain \cite{bethe}, \cite{yy} the delta function gases \cite{lieba}-\cite{yangb}, the 6--vertex model \cite{liebb},\cite{yangc}, the Hubbard model \cite{liebc}, and the 8--vertex model \cite{bax} for which the free energy (or ground state energy), and the excitation spectrum could be computed exactly. Since then it has been realized that a fundamental equation first seen by Onsager \cite{ons},\cite{wannier} in the Ising model and used in a profound way by Yang \cite{yangb} in the delta function gases and by Baxter \cite{bax} in the 8 vertex model could be used extend these computations to find many large classes of models for which free energies could be computed. These models which come from the Yang-Baxter (or star triangle equation) are what are now called the integrable models. The Ising model itself is the simplest case of such an integrable model. It thus seems to be a very natural conjecture which was made by Wu, myself and our collaborators the instant we made the discovery of the Painlev{\' e} representation of the Ising correlation function that there must be a similar representation for the correlation functions of all integrable models. To be more precise I mean the following {\bf Conjecture} {\it The correlation functions of integrable statistical mechanical models are characterized as the solutions of classically integrable equations (be they differential, integral or difference).} One major step in the advancement of this program was made by our next speaker, Alexander Zamolodchikov, who showed, with the invention of conformal field theory \cite{bpz}, that this conjecture is realized for models at the critical point. One of the major unsolved problems of integrable models today is to extend the linear equations which characterize correlation functions in conformal field theory to nonlinear equations for massive models. This will realize the goal of generalizing to all integrable models what we have learned for the correlation functions of the Ising model. This is an immense field of undertaking in which many people have and are making major contributions. It is surely not possible to come close to surveying this work in the few minutes left to me. I will therefore confine myself to a few remarks about things I have been personally involved with since completing the work with Wu in 1981 on the Ising model. \subsection{The chiral Potts model} In 1987 my coauthors H. Au-Yang, J.H.H. Perk, C.H. Sah, S. Tang, and M.L. Yan and I discovered \cite{ampty} the first example of an integrable model where, in technical language, the spectral variable lies on a curve of genus higher than one. This model has $N \geq 3$ states per site and is known as the integrable chiral Potts model. It is a particular case of a phenomenological model introduced for case $N=3$ in 1983 by Howes, Kadanoff and den Nijs \cite{hkd} in their famous study of level crossing transitions and is a generalization of the $N$ state model introduced by Von Gehlen and Rittenberg in 1985 \cite{gr} which generalizes Onsager's original solution of the Ising model \cite{ons},\cite{gst}. The Boltzmann weights were subsequently shown by Baxter, Perk and Au-Yang \cite{bpa} to have the following elegant form for $0\leq n \leq N-1$ \begin{equation} {W^h_{p,q}(n)\over W^h_{p,q}(0)}=\prod_{j=1}^{n}({d_pb_q-a_pc_q\omega^j \over b_pd_q-c_pa_q\omega^j}),~~ {W^v_{p,q}(n)\over W^v_{p,q}(0)}=\prod_{j=1}^n({\omega a_pd_q-d_pa_q\omega^j\over c_pb_q-b_pc_q\omega^j}) \label{cpw} \end{equation} where $\omega = e^{2 \pi i/N}.$ The variables $a_p,b_p,c_p,d_p$ and $a_q,b_q,c_q,d_q$ satisfy the equations \begin{equation} a^N+kb^N=k'd^N,~~~ka^N+b^N=k'c^N \label{curve} \end{equation} with $k^2+k'^2=1$ and this specifies a curve of genus $N^3-2N^2+1.$ When $N=2$ the Boltzmann weights reduce to those of the Ising model (\ref{ising}) with $H=0$ and the curve (\ref{curve}) reduces to the elliptic curve of genus 1. However when $N\geq 3$ the curve has genus higher than one. This is the first time that such higher genus curves has arisen in the Boltzmann weights of integrable models. This model is out of the class of all previously known models and raises a host of unsolved questions which are related to some of the most intractable problems of algebraic geometry which have been with us for 150 years. As an example of these new occurrences of ancient problems we can consider the spectrum of the transfer matrix \begin{equation} T_{\{l,l'\}}=\prod_{j=1}^{\cal N}W_{p,q}^v(l_j-l'_j)W_{p,q}^h(l_j-l'_{j+1}). \end{equation} This transfer matrix satisfies the commutation relation \begin{equation} [T(p,q),T(p,q')]=0 \end{equation} and also satisfies functional equations on the Riemann surface at points connected by the automorphism $R(a_q,b_q,c_q,d_q)=(b_q,\omega_q,d_q,c_q).$ For the Ising case $N=2$ this functional equation reduces to an equation which can be solved using elliptic theta functions. Most unhappily, however, for the higher genus case the analogous solution requires machinery from algebraic geometry which does not exist. For the problem of the free energy Baxter \cite{baxfree} has devised an ingenious method of solution which bypasses algebraic geometry completely but even here some problems remain in extending the method to the complete eigenvalue spectrum \cite{mr}. The problem is even more acute for the order parameter of the model. For the $N$ state models there are several order parameters parameterized by an integer index $n$ where $1\leq n\leq N-1.$ For these order parameters $M_n$ we conjectured \cite{al} 10 years ago from perturbation theory computations that \begin{equation} M_n=(1-k^2)^{n(N-n)/2N^2}. \end{equation} When $N=2$ this is exactly the result announced by Onsager \cite{onsb} in 1948 and proven by Yang \cite{yanga} for the Ising model in 1952. For the Ising model it took only three years to go from conjecture to proof. But for the chiral Potts model a decade has passed and even though Baxter \cite{baxe}--\cite{baxc} has produced several elegant formulations of the problem which all lead to the correct answer for the Ising case none of them contains enough information to solve the problem for $N\geq 3.$ In one approach \cite{baxe} the problem is reduced the the evaluation of a path ordered exponential of noncommuting variables on a Riemann surface. This sounds exactly like problems encountered in non Abelian gauge theory but, unfortunately, there is nothing in the field theory literature that helps. In another approach \cite{baxc} a major step in the solution involves the explicit reconstruction of a meromorphic function from a knowledge of its zeros and poles. This is a classic problem in algebraic geometry for which in fact no explicit answer is known either. Indeed the unsolved problems arising from the chiral Potts model are so resistant to all known mathematics that I have reduced my frustration to the following epigram; {\it The nineteenth century saw many brilliant creations of the human mind. Among them are algebraic geometry and Marxism. In the late twentieth century Marxism has been shown to be incapable of solving any practical problem but we still do not know about algebraic geometry.} It must be stressed again that the chiral Potts model was not invented because it was integrable but was found to be integrable after it was introduced to explain experimental data. In a very profound way physics is here far ahead of mathematics. \subsection{Exclusion statistics and Rogers-Ramanujan identities} One particularly important property of integrable systems is seen in the spectrum of excitations above the ground state. In all known cases these spectra are of the quasiparticle form in which the energies of multiparticle states are additively composed of single particle energies $e_{\alpha}(P)$ \begin{equation} E_{ex}-E_0=\sum_{\alpha=1}^n\sum_{j=1}^{m_\alpha}e_{\alpha}(P_j^{\alpha}) \end{equation} with the total momentum \begin{equation} P=\sum_{\alpha}^n\sum_{j=1}^{m_\alpha}P^{\alpha}_j~~({\rm mod}~2\pi). \end{equation} Here $n$ is the number of types of quasi-particles and there are $m_{\alpha}$ quasiparticles of type $\alpha.$ The momenta in the allowed states are quantized in units of $2\pi/M$ and are chosen from the sets \begin{equation} P^{\alpha}_j\in \{P_{\rm min}^{\alpha}({\bf m}), P_{\rm min}^{\alpha}({\bf m})+{2\pi\over M}, P_{\rm min}^{\alpha}({\bf m})+{4\pi\over M},\cdots, P_{\rm max}^{\alpha}({\bf m})\} \label{rules} \end{equation} with the Fermi exclusion rule \begin{equation} P_{j}^{\alpha}\neq P_{k}^{\alpha}~~{\rm for} j\neq k~~{\rm and~ all}~~ \alpha \label{fermi} \end{equation} and \begin{equation} P^{\alpha}_{\rm min}({\bf m})={\pi\over M}[({\bf m}({\bf B}-1))_{\alpha}-A_{\alpha}+1]~~{\rm and}~~ P^{\alpha}_{\rm max}=-P^{\alpha}_{\rm min} +{2\pi\over M}({{\bf u}\over 2}-{\bf A})_{\alpha} \end{equation} where if some $u_{\alpha}=\infty$ the corresponding $P_{\rm max}^{\alpha}=\infty.$ If some $e_{\alpha}(P)$ vanishes at some momentum (say 0) the system is massless and for $P\sim 0$ a typical behavior is $e_{\alpha}=v|P|$ where $v$ is variously called the speed of light or sound or the spin wave velocity. The important feature of the momentum selection rules (\ref{rules}) is that in addition to the fermionic exclusion rule (\ref{fermi}) is the exclusion of a certain number of momenta at the edge of the momentum zones which is proportional to the number of quasiparticles in the state. For the Ising model at zero field there is only one quasiparticle and $P_{\rm min}=0$ so the quasiparticle is exactly the same as a free fermion. However, for all other cases the $P_{\rm min}$ is not zero and exclusion does indeed take place. This is a very explicit characterization of the generalization which general integrable models make over the Ising model. The exclusion rules (\ref{rules}) lead to what have been called fractional (or exclusion) statistics by Haldane \cite{hal}. On the other hand they make a remarkable and beautiful connection with the mathematical theory of Rogers-Ramanujan identities and conformal field theory. We have found that these exclusion rules allow a reinterpretation of all conformal field theories (which are usually discussed in terms of a bosonic Fock space using a Feigin and Fuchs construction \cite{ff}) in terms of a set of fermionic quasiparticles \cite{kkmm}. What is most surprising is that there is not just one fermionic representation for each conformal field theory but there are at least as many distinct fermionic representations as there are integrable perturbations. The search for the complete set of fermionic representations is ongoing and I will only mention here that we have extensive results for the integrable perturbations of the minimal models $M(p,p')$ for the $\phi_{1,3}$ perturbation \cite{bms} and the $\phi_{2,1}$ and $\phi_{1,5}$ perturbations \cite{bmp}. \section{Beyond Integrability} There is one final problem of the hidden field of integrable models which I want to discuss. Namely the question of what is the relation of an integrable model to a generic physical system which does not satisfy a Yang-Baxter equation. For much of my career I have been told by many that these models are just mathematical curiosities which because they are integrable can, by that very fact, have nothing to do with real physics. But on the other hand the fact remains that all of the phenomenological insight we have into real physics phase transitions as embodied in the notions of critical exponents, scaling theory and universality which have served us well for 35 years either all come from integrable models or are all confirmed by the the solutions of integrable models. So if integrable models leave something out we have a very poor idea of what it is. Therefore it is greatly interesting that several months ago Bernie Nickel \cite{bn} sent around a preprint in which he made the most serious advance in the study of the Ising model susceptibility since our 1976 paper \cite{mccoyf}. In that paper in addition to the Painlev{\' e} representation of the two point function we derive an infinite series for the Ising model susceptibility where the $n^{th}$ term in the series involves an $n^{th}$ order integral. When the integrals in this expansion are scaled to the critical point each term contributes to the leading singularity of the susceptibility $|T-T_c|^{-7/4}$ However Nickel goes far beyond this scaling and for the isotropic case where $E^v=E^h=E$ in term of the variable $v=\sinh 2E/kT$ he shows that successive terms in the series contribute singularities that eventually become dense on the unit circle in the complex plane $|v|=1.$ From this he concludes that unless unexpected cancelations happen that there will be natural boundaries in the susceptibility on $|v|=1$. This would indeed be a new effect which could make integrable models different from generic models, Such natural boundaries have been suggested by several authors in the past including Guttmann \cite{gut}, and Orrick and myself \cite{mo} on the basis of perturbation studies of nonintegrable models which show ever increasingly complicated singularity structures as the order of perturbation increases; a complexity which magically disappears when an integrability condition is imposed. This connection between integrability and analyticity was first emphasized by Baxter \cite{baxd} long ago in 1980 when he emphasized that the Ising model in a magnetic field satisfies a functional equation very analogous to the zero field Ising model but that the Ising model in a field lacks the analyticity properties need for a solution. The proof of Nickel's conjecture will, if correct, open up a new view on what it means to be integrable. \section{Conclusion} I hope that I have conveyed to to you some of the excitement and challenges of the field of integrable models in statistical mechanics. The problems are physically important, experimentally accessible, and mathematically challenging. The field has been making constant progress since the first work of Bethe in 1931 and Onsager in 1944. So it might be thought that, even though the problems are hard, it would command the attention of some of the most powerful researchers in a large number of institutions. But as I indicated in the beginning of this talk this is in fact not the case. Most physics departments are more or less divided into the same divisions as is the APS. Thus it is quite typical to find departments with a condensed matter group, a nuclear physics group, and high energy group, an astrophysics group and an atomic and molecular group. But as I mentioned at the beginning, none of the work I have discussed in this talk fits naturally into these categories and thus if departments hire people in the mainstream of the existing divisions of the APS no one doing research in integrable models in statistical mechanics will ever be hired. So while I am deeply honored and grateful for the award of the 1999 Heineman prize for mathematical physics there is still another honor I am looking for. It is to receive a letter from the chairman of a physics department which reads as follows: {\it \noindent {Dear Prof. McCoy,} Thank you for the recommendation you recently made to us concerning the hiring of a new faculty member. We had not considered hiring anyone in the area of physics represented by your candidate, but after reading the resume and publications we decided that you were completely correct that the candidate is doing outstanding work which will bring an entirely new area of research to our department. We are very pleased to let you know that the university has made an offer of a faculty appointment to your candidate which has been accepted today. Thank you very much for your help and advice.} I have actually received one such letter in my life. If I am fortunate I hope to receive a few more before the end of my career. The 21st century is long and anything is still possible. \vspace{.5in} {\bf Acknowledgments} This work is supported in part by the National Science Foundation under grant DMR 97-03543. \vspace{.5in}
\section{Introduction} Compton scattering of real photons (RCS) is one of the simplest reactions for obtaining information on the structure of a stable composite system. When expanded in the frequency of the photon, the leading-order term of the low-energy scattering amplitude is specif\/ied by the model-independent Thomson limit in terms of the charge and the mass of the target. Genuine structure effects f\/irst appear at second order and can be parametrized in terms of the electric and magnetic polarizabilities (for an overview see, e.g., Refs.\ \cite{Holstein_90,Lvov_93,Scherer_99}). As there is no stable pion target, the empirical information on the electromagnetic polarizabilities has been extracted from high-energy pion-nucleus bremsstrahlung \cite{Antipov_83,Antipov_85} and radiative pion photoproduction off the nucleon \cite{Aibergenov_86}. In principle, the electromagnetic polarizabilities of the pion also enter into the crossed process $\gamma\gamma\to\pi\pi$. However, there is some debate concerning the accuracy of extracting these quantities from the crossed channel \cite{Babusci_92,Kaloshin_92,Donoghue_93,Kaloshin_94,Portoles_95}. From a theoretical point of view, a precise determination of the pion polarizabilities is of great importance, since (approximate) chiral symmetry allows one to predict the electromagnetic polarizabilities of the charged pion in terms of the radiative decay $\pi^+\to e^+\nu_e\gamma$ \cite{Terentev_73}. Corrections to the leading-order PCAC result have been calculated at ${\cal O}(p^6)$ in chiral perturbation theory and turn out to be be rather small \cite{Buergi_96}. New experiments are presently being carried out \cite{Ahrens_95} or have been proposed \cite{Moinester_97,Gorringe_98} to signif\/icantly reduce the uncertainties in the empirical results and thus subject the predictions of chiral symmetry to a stringent test. Clearly, the possibilities to investigate the structure of the target increase substantially if virtual photons are used, because energy and three-momentum can be varied independently and, furthermore, the longitudinal component of the transition current can be explored. In particular, virtual Compton scattering (VCS) off the nucleon, as tested in the reaction $e^-+p\to e^- + p +\gamma$, has attracted considerable interest (see, e.g., Refs.\ \cite{dHose_97,Guichon_98}). The pion-VCS amplitude of $\gamma^\ast+\pi\to \gamma+\pi$ can, in principle, be studied through the inelastic scattering of high-energy pions off atomic electrons, $\pi+e^-\to\pi+e^-+\gamma$. Such events are presently analyzed as part of the SELEX E781 experiment \cite{Moinester_99}. In this paper, we will investigate the VCS reaction $\gamma^\ast+\pi\to \gamma+\pi$ in the framework of chiral perturbation theory at ${\cal O}(p^4)$. We will f\/irst give a short survey of chiral perturbation theory and then def\/ine our conventions for the VCS invariant amplitude. We then discuss the result for the soft-photon and residual amplitudes, respectively. Finally, the model-dependent residual amplitude is analyzed in terms of alternative def\/initions of generalized polarizabilities. \section{The chiral Lagrangian} Chiral perturbation theory (ChPT) \cite{Weinberg_79,Gasser_84,Gasser_85} is based on the chiral $\mbox{SU(2)}_L\times \mbox{SU(2)}_R$ symmetry of QCD in the limit of vanishing $u$- and $d$-quark masses. The assumption of spontaneous symmetry breaking down to $\mbox{SU(2)}_V$ gives rise to three massless pseudoscalar Goldstone bosons with vanishing interactions in the limit of zero energies. These Goldstone bosons are identif\/ied with the physical pion triplet, the nonzero pion masses resulting from an explicit symmetry breaking in QCD through the quark masses. The effective Lagrangian of the pion interaction is organized in a so-called momentum expansion, \begin{equation} {\cal L}_{\mbox{\footnotesize eff}} ={\cal L}_2+{\cal L}_4+\cdots, \end{equation} where the subscripts refer to the order in the expansion. Interactions with external f\/ields, such as the electromagnetic f\/ield, as well as explicit symmetry breaking due to the f\/inite quark masses, are systematically incorporated into the effective Lagrangian. Covariant derivatives and quark-mass terms count as ${\cal O}(p)$ and ${\cal O}(p^2)$, respectively. Weinberg's power counting scheme \cite{Weinberg_79} allows for a classif\/ication of the Feynman diagrams by establishing a relation between the momentum expansion and the loop expansion. The most general chiral Lagrangian at ${\cal O}(p^2)$ is given by \begin{equation} \label{l2} {\cal L}_2 = \frac{F^2}{4} \mbox{Tr} \left[ D_{\mu} U (D^{\mu}U)^{\dagger} +\chi U^{\dagger}+ U \chi^{\dagger} \right], \end{equation} where $U$ is a unimodular unitary $(2\times 2)$ matrix, transforming as $V_R U V_L^\dagger$ for $(V_L,V_R)\in \mbox{SU(2)}_L\times\mbox{SU(2)}_R$. As a parametrization of $U$ we will use \begin{equation} \label{paru} U(x)=\frac{\sigma(x)+i\vec{\tau}\cdot\vec{\pi}(x)}{F}, \quad \sigma^2(x)+\vec{\pi}^2(x)=F^2, \end{equation} where $F$ denotes the pion-decay constant in the chiral limit: $F_\pi=F[1+{\cal O}(\hat{m})]=92.4$ MeV. We will work in the isospin-symmetric limit $m_u=m_d=\hat{m}$. The quark mass is contained in $\chi=2 B_0 \hat{m}=m^2_\pi$ at ${\cal O}(p^2)$, where $B_0$ is related to the quark condensate $<\!\!\bar{q}q\!\!>$. The covariant derivative $D_\mu U = \partial_\mu U +\frac{i}{2}e A_\mu [\tau_3,U]$ contains the coupling to the electromagnetic f\/ield $A_\mu$. The most general structure of ${\cal L}_4$, f\/irst obtained by Gasser and Leutwyler (see Eq.\ (5.5) of Ref.\ \cite{Gasser_84}), reads, in the standard trace notation, \begin{eqnarray} \label{l4gl} {\cal L}^{GL}_4 &=& \frac{l_1}{4} \left\{\mbox{Tr}[D_{\mu}U (D^{\mu}U)^{\dagger}] \right\}^2 +\frac{l_2}{4}\mbox{Tr}[D_{\mu}U (D_{\nu}U)^{\dagger}] \mbox{Tr}[D^{\mu}U (D^{\nu}U)^{\dagger}] +\frac{l_3}{16}\left[\mbox{Tr}(\chi U^\dagger+ U\chi^\dagger)\right]^2 \nonumber\\ &&+\frac{l_4}{4}\mbox{Tr}[D_\mu U(D^\mu\chi)^\dagger +D_\mu\chi(D^\mu U)^\dagger] +l_5\left[\mbox{Tr}(F^R_{\mu\nu}U F^{\mu\nu}_LU^\dagger) -\frac{1}{2}\mbox{Tr}(F_{\mu\nu}^L F^{\mu\nu}_L +F_{\mu\nu}^R F^{\mu\nu}_R)\right]\nonumber\\ &&+i\frac{l_6}{2}\mbox{Tr}[ F^R_{\mu\nu} D^{\mu} U (D^{\nu} U)^{\dagger} + F^L_{\mu\nu} (D^{\mu} U)^{\dagger} D^{\nu} U] -\frac{l_7}{16}\left[\mbox{Tr}(\chi U^\dagger-U\chi^\dagger)\right]^2 +\cdots, \end{eqnarray} where three terms containing only external f\/ields have been omitted. For the electromagnetic interaction, the f\/ield-strength tensors are given by $F^{\mu\nu}_L=F^{\mu\nu}_R=-\frac{e}{2}\tau_3 (\partial^\mu A^\nu-\partial^\nu A^\mu)$. \section{Conventions} In the following, we will discuss the VCS amplitude for $\gamma^\ast(q,\epsilon)+\pi^i(p_i)\to\gamma(q',\epsilon')+\pi^j(p_f)$ ($q^2\leq 0$, $q'^2=0$, $q'\cdot\epsilon'=0$). Throughout the calculation we use the conventions of Bjorken and Drell \cite{Bjorken_1964} with $e^2/4\pi\approx1/137$, $e>0$. For the isospin decomposition of the invariant amplitude we use \begin{equation} \label{isospin} {\cal M}_{ij}=\delta_{ij}{\cal A}+(\delta_{ij}-\delta_{i3}\delta_{j3}) {\cal B}, \end{equation} where $i$ and $j$ denote the {\em cartesian} isospin indices of the initial and f\/inal pions, respectively. With the def\/inition \begin{displaymath} |\pi^\pm(p)\!>=\frac{1}{\sqrt{2}}[a_1^\dagger(p)\pm i a_2^\dagger(p)]|0\!>, \quad |\pi^0(p)\!>=a_3^\dagger(p)|0\!>, \end{displaymath} we may express the physical amplitudes in terms of the isospin amplitudes \begin{eqnarray} \label{mcharged} {\cal M}_{\pi^+}={\cal M}_{\pi^-}&=&\frac{1}{2}({\cal M}_{11}+{\cal M}_{22}) ={\cal A}+{\cal B},\\ \label{mneutral} {\cal M}_{\pi^0}&=&{\cal M}_{33}={\cal A}. \end{eqnarray} We split the contributions to ${\cal M}_{ij}$ into a pole piece ($P$) and a one-particle-irreducible, residual part ($R$), ${\cal A}={\cal A}_P+{\cal A}_R$, ${\cal B}={\cal B}_P+{\cal B}_R$ (see Fig.\ \ref{vcsdiagrams.fig}). Since the $\pi^0$ is its own antiparticle, the electromagnetic vertex $\pi^0\pi^0\gamma^\ast$ vanishes due to charge-conjugation invariance and hence ${\cal A}_P\equiv 0$. In general, the pole piece ${\cal B}_P$ and the one-particle-irreducible piece ${\cal B}_R$ are not separately gauge invariant. \section{Soft-photon amplitude} According to Weinberg's power counting, a calculation of the $s$- and $u$-channel pole terms at ${\cal O}(p^4)$ involves the renormalized irreducible vertex at ${\cal O}(p^4)$, \begin{equation} \label{gammamu} \Gamma^\mu(p',p)=(p'+p)^\mu F(q^2)+(p'-p)^\mu \frac{p'^2-p^2}{q^2}[ 1-F(q^2)],\quad q=p'-p, \end{equation} where $F(q^2)$ is the prediction for the electromagnetic form factor of the pion (see Eq.\ (15.3) of Ref.\ \cite{Gasser_84}). To that order, the renormalized propagator is simply given by \begin{equation} \label{prop} i\Delta_R(p)=\frac{i}{p^2-m_\pi^2+i0^+}, \end{equation} with $m^2_\pi$ the ${\cal O}(p^4)$ result for the pion mass squared (see Eq.\ (12.2) of Ref.\ \cite{Gasser_84}). Note that Eqs.\ (\ref{gammamu}) and (\ref{prop}) satisfy the Ward-Takahashi identity \cite{Ward_50,Takahashi_57} $q_\mu \Gamma^\mu(p',p)=\Delta_R^{-1}(p')-\Delta_R^{-1}(p)$. With these ingredients the result for ${\cal B}_P$ at ${\cal O}(p^4)$ reads \begin{equation} \label{bp} {\cal B}_P=-ie^2\left\{F(q^2)\left[\frac{2p_f\cdot\epsilon'^\ast\, (2p_i+q)\cdot\epsilon}{s-m_\pi^2} +\frac{(2p_f-q)\cdot\epsilon\, 2p_i\cdot\epsilon'^\ast}{u-m_\pi^2}\right] +2 q\cdot\epsilon\, q\cdot\epsilon'^\ast \frac{1-F(q^2)}{q^2}\right\}, \end{equation} which is easily seen not to be gauge invariant by itself. The set of one-particle-irreducible diagrams is shown in Fig.\ \ref{classb.fig} and gives rise to a residual part of the form \begin{equation} \label{br} {\cal B}_R=ie^2\left[2\epsilon\cdot\epsilon'^\ast +2(q^2\epsilon\cdot\epsilon'^\ast-q\cdot \epsilon\, q\cdot\epsilon'^\ast)\frac{F(q^2)-1}{q^2}\right]+\tilde{\cal B}_R. \end{equation} We combine Eqs.\ (\ref{bp}) and (\ref{br}) into the form \begin{equation} \label{b} {\cal B}=\tilde{\cal B}_P+\tilde{\cal B}_R, \end{equation} where \begin{equation} \label{tildebp} \tilde{\cal B}_P=-ie^2 F(q^2)\left[\frac{2p_f\cdot\epsilon'^\ast\, (2p_i+q)\cdot\epsilon}{s-m_\pi^2} +\frac{(2p_f-q)\cdot\epsilon\, 2p_i\cdot\epsilon'^\ast}{u-m_\pi^2} -2\epsilon\cdot\epsilon'^\ast\right], \end{equation} with the result that $\tilde{\cal B}_P$ and $\tilde{\cal B}_R$ are now separately gauge invariant. In particular, $\tilde{\cal B}_P$ has the form of the soft-photon result obtained in Eq.\ (10) of Ref.\ \cite{Fearing_98}. A somewhat different approach for obtaining the soft-photon result can be found in Ref.\ \cite{Lvov_99}. \section{Residual amplitudes} As has been discussed in detail in Ref.\ \cite{Drechsel_1997}, a gauge-invariant parametrization of the residual amplitude for $\gamma^\ast+\pi\to \gamma+\pi$ can be written in terms of three invariant functions $f_i(q^2,q\cdot q',q\cdot P)$, where $P=p_i+p_f$. At ${\cal O}(p^4)$, the result for the residual isospin amplitudes ${\cal A}_R$ and $\tilde{\cal B}_R$ reads: \begin{eqnarray} \label{ar} {\cal A}_R&=&-ie^2(q'\cdot\epsilon\, q\cdot \epsilon'^\ast-q\cdot q' \epsilon\cdot\epsilon'^\ast) \frac{m^2_\pi+2q\cdot q'-q^2}{8\pi^2F_\pi^2 q\cdot q'} {\cal G}(q^2,q\cdot q'),\\ \label{tbr} \tilde{\cal B}_R&=&-ie^2(q'\cdot\epsilon \,q\cdot \epsilon'^\ast-q\cdot q' \epsilon\cdot\epsilon'^\ast) \left[-\frac{4(2 l_5^r-l_6^r)}{F^2_\pi} -\frac{2 m^2_\pi+2q\cdot q'-q^2}{16\pi^2 F^2_\pi q\cdot q'} {\cal G}(q^2,q\cdot q')\right], \end{eqnarray} where the combination $2l^r_5-l^r_6=(2.85\pm 0.42)\times 10^{-3}$ is determined through the decay $\pi^+\to e^+\nu_e\gamma$. In Eqs.\ (\ref{ar}) and (\ref{tbr}) we have introduced the abbreviation \begin{equation} \label{functiong} {\cal G}(q^2,q\cdot q')= 1+\frac{m^2_\pi}{q\cdot q'}\left[J^{(-1)}(a)-J^{(-1)}(b)\right] -\frac{q^2}{2q\cdot q'}\left[J^{(0)}(a)-J^{(0)}(b)\right], \end{equation} where \begin{displaymath} J^{(n)}(x):=\int_0^1 dy y^n\ln[1+x(y^2-y)-i0^+] \end{displaymath} and \begin{displaymath} a:=\frac{q^2}{m^2_\pi},\quad b:=\frac{q^2-2q\cdot q'}{m^2_\pi}. \end{displaymath} The one-loop integrals $J^{(0)}$ and $J^{(-1)}$ are given by (see Appendix C of Ref.\ \cite{Bellucci_1994}\footnote{In reproducing these results we found Refs.\ \cite{Barbieri_1972} and \cite{tHooft_1979} useful.}) \begin{eqnarray*} J^{(0)}(x) &=& \left \{ \begin{array}{l} -2-\sigma\ln\left(\frac{\sigma-1}{\sigma+1}\right)\quad(x<0),\\ -2+2\sqrt{\frac{4}{x}-1}\,\mbox{arccot} \left(\sqrt{\frac{4}{x}-1}\right)\quad (0\le x<4),\\ -2-\sigma\ln\left(\frac{1-\sigma}{1+\sigma}\right)-i\pi\sigma \quad(4<x), \end{array} \right.\\ J^{(-1)}(x)&=&\left \{ \begin{array}{l} \frac{1}{2}\ln^2\left(\frac{\sigma-1}{\sigma+1}\right)\quad(x< 0),\\ -\frac{1}{2}\arccos^2\left(1-\frac{x}{2}\right)\quad (0\le x <4),\\ \frac{1}{2}\ln^2\left(\frac{1-\sigma}{1+\sigma}\right)-\frac{\pi^2}{2} +i\pi\ln\left(\frac{1-\sigma}{1+\sigma}\right)\quad (4< x), \end{array} \right. \end{eqnarray*} with \begin{displaymath} \sigma(x)=\sqrt{1-\frac{4}{x}},\quad x\notin [0,4]. \end{displaymath} A comparison of Eqs.\ (\ref{ar}) and (\ref{tbr}) with Eq.\ (18) of Ref.\ \cite{Drechsel_1997} shows that, at ${\cal O}(p^4)$, only one of the three functions $f_i(q^2,q\cdot q',q\cdot P)$ contributes, i.e., $f_2=f_3=0$. Furthermore, at this order in the chiral expansion, the function $f_1$ does not depend on $q\cdot P=q'\cdot P$. Our result for ${\cal A}_R$ is in agreement with Ref.\ \cite{Belkov_97}, where the photoproduction of neutral pion pairs in the Coulomb f\/ield of a nucleus was studied. \section{Generalized polarizabilities of Guichon, Liu, and Thomas} In order to discuss the generalized polarizabilities, we expand the function ${\cal G}$ of Eq.\ (\ref{functiong}) for negative $q^2$ around $q\cdot q'=0$, \begin{equation} \label{functiongexp} {\cal G}(q^2,q\cdot q')=-\frac{q\cdot q'}{m^2_\pi} J^{(0)'}\left(\frac{q^2}{m^2_\pi}\right)+ {\cal O}[(q\cdot q')^2],\quad J^{(0)'}(x)=\frac{d J^{(0)}(x)}{dx}, \end{equation} where \begin{equation} \label{jop} J^{(0)'}(x)=\frac{1}{x}\left[1-\frac{2}{x\sigma} \ln\left(\frac{\sigma-1}{\sigma+1}\right)\right] =\frac{1}{x}\left[1+2J^{(-1)'}(x)\right],\quad x< 0. \end{equation} For the charged and neutral pion we obtain, respectively, \begin{eqnarray} \label{f1pp} f_1^{\pi^\pm}(q^2,q\cdot q',q\cdot P)&=&-\frac{4(2l^r_5-l_6^r)}{F^2_\pi} +\frac{2q\cdot q'-q^2}{16\pi^2 F^2_\pi q\cdot q'} {\cal G}(q^2,q\cdot q')\nonumber\\ &=&-\frac{4(2l^r_5-l_6^r)}{F^2_\pi}+\frac{q^2}{16\pi^2 F^2_\pi m^2_\pi} J^{(0)'}\left(\frac{q^2}{m^2_\pi}\right) +{\cal O}(q\cdot q'),\\ \label{f1p0} f_1^{\pi^0}(q^2,q\cdot q',q\cdot P)&=& -\frac{1}{8\pi^2 F^2_\pi}\left(1-\frac{q^2}{m^2_\pi}\right) J^{(0)'}\left(\frac{q^2}{m^2_\pi}\right) +{\cal O}(q\cdot q'). \end{eqnarray} We will f\/irst discuss the generalized polarizabilities as def\/ined in Ref.\ \cite{Guichon_1995}, where the residual amplitude was analyzed in the photon-pion center-of-mass frame in terms of a multipole expansion. Only terms {\em linear} in the frequency of the f\/inal photon were kept, and the result was parametrized in terms of ``generalized polarizabilities.'' The connection with the covariant approach was established in Ref.\ \cite{Drechsel_1997}, where it was also found that only two of the three polarizabilities $P^{(01,01)0}$, $P^{(11,11)0}$, and $\hat{P}^{(01,1)0}$ of Ref.\ \cite{Guichon_1995} are independent, once the constraints due to charge conjugation are combined with particle-crossing symmetry. According to Eqs.\ (35) and (36) of Ref.\ \cite{Drechsel_1997} we def\/ine generalized electric and magnetic polarizabilities $\alpha(|\vec{q}|^2)$ and $\beta(|\vec{q}|^2)$, respectively, as \begin{eqnarray} \label{alpha1} \alpha(|\vec{q}|^2)&\equiv& -\frac{e^2}{4\pi}\sqrt{\frac{3}{2}}P^{(01,01)0}(|\vec{q}|)\nonumber\\ &=& \frac{e^2}{8\pi m_\pi}\sqrt{\frac{m_\pi}{E_i}} \left[-f_1(\omega_0^2-|\vec{q}|^2,0,0)+2m_\pi\frac{|\vec{q}|^2}{\omega_0} f_2(\omega_0^2-|\vec{q}|^2,0,0)\right],\\ \label{beta1} \beta(|\vec{q}|^2)&\equiv&-\frac{e^2}{4\pi}\sqrt{\frac{3}{8}}P^{(11,11)0} (|\vec{q}|)= \frac{e^2}{8\pi m_\pi}\sqrt{\frac{m_\pi}{E_i}} f_1(\omega_0^2-|\vec{q}|^2,0,0), \end{eqnarray} where $\omega_0=q_0|_{\omega'=0}=m_\pi-\sqrt{m_\pi^2+|\vec{q}|^2}$. A few remarks are in order at this point. \begin{enumerate} \item In our present work, we strictly stick to the convention of Ref.\ \cite{Bjorken_1964}. This is why Eqs.\ (\ref{alpha1}) and (\ref{beta1}) differ by an overall factor $1/2m_\pi$ from Ref.\ \cite{Drechsel_1997}, where in Eq.\ (1) an additional factor $2m_\pi$ was introduced for the spin-0 case. \item The variable $q^2$ only appears in the combination $q^2/m^2_\pi$, resulting in \begin{displaymath} \left.\frac{q^2}{m^2_\pi}\right|_{\omega'=0}=2\frac{m_\pi-E_i}{m_\pi},\quad E_i=\sqrt{m_\pi^2+|\vec{q}|^2}. \end{displaymath} \item The factor $\sqrt{m_\pi/E_i}$ originates from an additional normalization factor ${\cal N}$ in Eq.\ (32) of Ref.\ \cite{Guichon_1995}, such that \begin{displaymath} \frac{2m_\pi}{\sqrt{4 E_i E_f}}\stackrel{\omega'\to 0}{\to} \sqrt{\frac{m_\pi}{E_i}}. \end{displaymath} \end{enumerate} Using the results of Eqs.\ (\ref{f1pp}) and (\ref{f1p0}) together with $f_2=0$, we then obtain \begin{eqnarray} \label{alphapp1} \alpha_{\pi^\pm}(|\vec{q}|^2)&=&-\beta_{\pi^\pm}(|\vec{q}|^2) = \frac{e^2}{8\pi m_\pi}\sqrt{\frac{m_\pi}{E_i}} \left[\frac{4(2l^r_5-l^r_6)}{F^2_\pi}-2 \frac{m_\pi-E_i}{m_\pi} \frac{1}{(4\pi F_\pi)^2} J^{(0)'}\left(2\frac{m_\pi-E_i}{m_\pi}\right)\right], \nonumber\\ &&\\ \label{alphap01} \alpha_{\pi^0}(|\vec{q}|^2)&=&-\beta_{\pi^0}(|\vec{q}|^2) =\frac{e^2}{4\pi} \frac{1}{(4\pi F_\pi)^2 m_\pi}\sqrt{\frac{m_\pi}{E_i}} \left(1-2\frac{m_\pi- E_i}{m_\pi}\right) J^{(0)'}\left(2\frac{m_\pi- E_i}{m_\pi}\right). \end{eqnarray} At the one-loop level, the $|\vec{q}|^2$ dependence is entirely given in terms of the pion mass $m_\pi$ and the pion-decay constant $F_\pi$, i.e., no additional ${\cal O}(p^4)$ low-energy constant enters. At $|\vec{q}|^2=0$, Eqs.\ (\ref{alphapp1}) and (\ref{alphap01}) reduce to the RCS polarizabilities \cite{Holstein_90} \begin{eqnarray} \bar{\alpha}_{\pi^\pm}&=&-\bar{\beta}_{\pi^\pm}=\frac{e^2}{4\pi} \frac{2}{m_\pi F^2_\pi}(2l^r_5-l^r_6) =(2.68\pm0.42)\times 10^{-4}\,\mbox{fm}^3,\\ \bar{\alpha}_{\pi^0}&=&-\bar{\beta}_{\pi^0} =-\frac{e^2}{4\pi}\frac{1}{96\pi^2 F^2_\pi m_\pi} =-0.50 \times 10^{-4}\,\mbox{fm}^3, \end{eqnarray} where we made use of $J^{(0)'}(0)=-\frac{1}{6}$. At ${\cal O}(p^6)$, the RCS predictions for the charged pion read $\bar{\alpha}_{\pi\pm}=(2.4\pm 0.5)\times 10^{-4}\,\mbox{fm}^3$ and $\bar{\beta}_{\pi^\pm}=(-2.1\pm 0.5)\times 10^{-4}\,\mbox{fm}^3$ \cite{Buergi_96}. The corresponding corrections amount to a 12\% (24\%) change of the ${\cal O}(p^4)$ result, indicating a good convergence. We also note that the original degeneracy $\bar{\alpha}=-\bar{\beta}$ is lifted at ${\cal O}(p^6)$. The predictions of ChPT have to be compared with the empirical results $\bar{\alpha}_{\pi\pm}=(6.8\pm 1.4)\times 10^{-4}\,\mbox{fm}^3$ \cite{Antipov_83}, $\bar{\alpha}_{\pi\pm}=(20\pm 12)\times 10^{-4}\,\mbox{fm}^3$ \cite{Aibergenov_86}, and $\bar{\beta}_{\pi^\pm}=(-7.1\pm 4.6)\times 10^{-4}\,\mbox{fm}^3$ \cite{Antipov_85}. Clearly, an improved accuracy is required to test the chiral predictions. For the neutral pion, the ${\cal O}(p^6)$ corrections turn out to be much larger, $\bar{\alpha}_{\pi^0}=(-0.35\pm 0.10)\times 10^{-4}\,\mbox{fm}^3$ and $\bar{\beta}_{\pi^0}=(1.50\pm 0.20)\times 10^{-4}\,\mbox{fm}^3$ \cite{Bellucci_1994}. \section{Alternative def\/inition of the generalized dipole polarizabilities} Another generalization of the RCS polarizabilities is obtained by parametrizing the invariant amplitude as\footnote{ A detailed discussion will be given in Ref.\ \cite{Lvov_99}.} \begin{equation} \label{mvcs2} -i{\cal M}=B_1 F^{\mu\nu} F'_{\mu\nu} +\frac{1}{4}B_2(P_\mu F^{\mu\nu})(P^\rho F'_{\rho\nu}) +\frac{1}{4}B_5(P^\nu q^\mu F_{\mu\nu})(P^\sigma q^\rho F'_{\rho\sigma}), \end{equation} where $F^{\mu\nu}$ and $F'_{\mu\nu}$ refer, respectively, to the gauge-invariant combinations \begin{displaymath} F^{\mu\nu}=-i q^\mu \epsilon^\nu+iq^\nu \epsilon^\mu,\quad F'_{\mu\nu}=iq'_\mu\epsilon'^\ast_\nu-iq'_\nu\epsilon'^\ast_\mu. \end{displaymath} The functions $B_1$, $B_2$, and $B_5$ are {\em even} functions of $P$. Introducing the suggestive notation $$\vec{E}=i(q_0\vec{\epsilon}-\vec{q}\epsilon_0),\quad \vec{B}=i\vec{q}\times\vec{\epsilon},\quad \vec{E}'=-i(q_0'\vec{\epsilon}\,'^\ast-\vec{q}\,'\epsilon_0'^\ast),\quad \vec{B}'=-i\vec{q}\,'\times\vec{\epsilon}\,'^\ast, $$ the structures of Eq.\ (\ref{mvcs2}) are particularly simple when evaluated in the pion Breit frame (p.B.f.) def\/ined by $\vec{P}=0$, \begin{eqnarray*} F^{\mu\nu} F'_{\mu\nu}&=&[-2\vec{E}\cdot\vec{E}'+2\vec{B}\cdot\vec{B}']_{ p.B.f.},\\ P_\mu F^{\mu\nu} P^\rho F'_{\rho\nu}&=& [-P^2_0 \vec{E}\cdot\vec{E}']_{p.B.f.},\\ P^\nu q^\mu F_{\mu\nu} P^\rho q^\sigma F'_{\sigma\rho}&=& [P^2_0 \vec{q}\cdot \vec{E}\vec{q}\cdot\vec{E}']_{p.B.f.}. \end{eqnarray*} Note that by def\/inition $[P^2_0]_{p.B.f.}=P^2$. In the p.B.f., Eq.\ (\ref{mvcs2}) can thus be expressed as \begin{equation} \label{mvcsnbf} -i{\cal M}=\left[2 B_1 \vec{B}\cdot\vec{B}' -\left(2B_1+\frac{P^2}{4} B_2\right) \vec{E}\cdot\vec{E}' +\frac{P^2}{4}B_5 \vec{q}\cdot \vec{E} \vec{q}\cdot\vec{E}'\right]_{p.B.f.}. \end{equation} Since $\vec{E}=\vec{E}_T+\vec{E}_L$, $\vec{E}\cdot\vec{E}'$ contains both transverse and longitudinal components with respect to $\hat{q}$, for which reason we will introduce the quantities $\alpha_T$ and $\alpha_L$ below: \begin{equation} \label{mvcsnbf2} -i{\cal M}=\left\{ 2 B_1 \vec{B}\cdot\vec{B}' -\left(2B_1+\frac{P^2}{4} B_2\right) \vec{E}_T\cdot\vec{E}' +\left[\frac{P^2}{4}B_5 |\vec{q}|^2-\left(2B_1+\frac{P^2}{4} B_2\right)\right] \vec{E}_L\cdot\vec{E}'\right\}_{p.B.f.}. \end{equation} We now consider the limit $\omega'\to 0$ of the residual amplitudes, for which $B_i^r\to b_i^r(q^2)$, and def\/ine three generalized dipole polarizabilities in terms of the invariants of Eq.\ (\ref{mvcs2}), \begin{eqnarray} \label{beta} 8\pi m_\pi\beta(q^2)&\equiv&2b_1^r(q^2),\\ \label{alphat} 8\pi m_\pi \alpha_T(q^2)&\equiv&-2b_1^r(q^2)-\left(M^2-\frac{q^2}{4}\right)b_2^r(q^2),\\ \label{alphal} 8\pi m_\pi\alpha_L(q^2)&\equiv&-2b_1^r(q^2)-\left(M^2-\frac{q^2}{4} \right)[b_2^r(q^2)+q^2 b_5^r(q^2)], \end{eqnarray} the superscript $r$ referring to the residual amplitudes beyond the soft-photon result. In general, the transverse and longitudinal electric polarizabilities $\alpha_T$ and $\alpha_L$ will differ by a term, vanishing however in the RCS limit $q^2=0$. Comparing with Eq.\ (\ref{mvcsnbf2}), the generalized dipole polarizabilities are seen to be def\/ined such that they multiply the structures $\vec{B}\cdot\vec{B}'$, $\vec{E}_T\cdot\vec{E}'$, and $\vec{E}_L\cdot\vec{E}'$, respectively, as $\omega'\to 0$. We note that $[\vec{B}\cdot\vec{B}']_{p.B.f.}$ and $ [\vec{E}_L\cdot \vec{E}']_{p.B.f.}$ are of ${\cal O}(\omega')$ whereas $[\vec{E}_T\cdot\vec{E}']_{p.B.f.}= {\cal O}(\omega'^2)$, i.e., that different powers of $\omega'$ have been kept. At $q^2=0$, the usual RCS polarizabilities are recovered, \begin{equation} \label{rcslimit} \beta(0)=\bar{\beta},\quad \alpha_L(0)=\alpha_T(0)=\bar{\alpha}. \end{equation} The connection to the generalized polarizabilities of Guichon {\em et al.} \cite{Guichon_1995} can either be established by direct comparison or via the results of Ref.\ \cite{Drechsel_1997}, \begin{eqnarray} \alpha(|\vec{q}|^2)&=&-\frac{e^2}{4\pi} \sqrt{\frac{3}{2}}P^{(01,01)0}( |\vec{q}|) =\sqrt{\frac{m_\pi}{E_i}}\alpha_L(\omega_0^2-\vec{q}\,^2),\\ \beta(|\vec{q}|^2)&=&-\frac{e^2}{4\pi} \sqrt{\frac{3}{8}}P^{(11,11)0}( |\vec{q}|) =\sqrt{\frac{m_\pi}{E_i}}\beta(\omega_0^2-\vec{q}\,^2), \end{eqnarray} with $\omega_0=q_0|_{\omega'=0}=m_\pi-E_i$, $\omega_0^2-\vec{q}\,^2=2m_\pi(m_\pi-E_i)$, and $E_i= \sqrt{m^2_\pi+\vec{q}\,^2}$, and all variables referring to the cm frame. Using Eqs.\ (35) - (37) of Ref.\ \cite{Drechsel_1997}, we f\/ind that the transverse electric dipole polarizability is part of a second-order contribution in $\omega'$ beyond the approximation of Guichon {\em et al.}, \begin{equation} \alpha_T(q^2)=\alpha_L(q^2)+\frac{e^2}{4\pi}(4M^2-q^2)q^2 \tilde{f}_3(q^2,0,0), \end{equation} where $q\cdot P \tilde{f}_3\equiv f_3$. At ${\cal O}(p^4)$, $f_2=f_3=0$, with the result of particularly simple expressions for the generalized dipole polarizabilities, \begin{eqnarray} \label{alphapp2} \alpha^{\pi^\pm}_L(q^2)&=& \alpha^{\pi^\pm}_T(q^2)= -\beta^{\pi^\pm}(q^2) = \frac{e^2}{8\pi m_\pi} \left[\frac{4(2l^r_5-l^r_6)}{F^2_\pi}-\frac{q^2}{m_\pi^2} \frac{1}{(4\pi F_\pi)^2} J^{(0)'}\left(\frac{q^2}{m_\pi^2}\right)\right], \\ \label{alphap02} \alpha^{\pi^0}_L(q^2)&=& \alpha^{\pi^0}_T(q^2)= -\beta^{\pi^0}(q^2) =\frac{e^2}{4\pi}\frac{1}{(4\pi F_\pi)^2m_\pi} \left(1-\frac{q^2}{m^2_\pi}\right) J^{(0)'}\left(\frac{q^2}{m^2_\pi}\right). \end{eqnarray} The results for the generalized dipole polarizabilities are shown in F\/ig.\ \ref{alpha.fig}. Even though chiral perturbation theory is only applicable for small external momenta, for the sake of completeness we also quote the asymptotic behavior as $q^2\to-\infty$, \begin{eqnarray} \label{asbpp} \alpha^{\pi^\pm}_L(q^2) &\to& \bar{\alpha}_{\pi^\pm}+3\bar{\alpha}_{\pi^0} =1.18\times 10^{-4}\,\mbox{fm}^3,\\ \alpha^{\pi^0}_L(q^2)&\to& 6 \bar{\alpha}_{\pi^0} =-3.0\times 10^{-4}\,\mbox{fm}^3. \end{eqnarray} As in the case of real Compton scattering, we expect the degeneracy $\alpha_L(q^2)=\alpha_T(q^2)=-\beta(q^2)$ to be lifted at the two-loop level. \section{Summary} We have calculated the invariant amplitudes for virtual Compton scattering off the pion, $\gamma^\ast+\pi\to \gamma +\pi$, at the one-loop level, ${\cal O}(p^4)$, in chiral perturbation theory. For the charged pion, the result may be decomposed into a gauge-invariant soft-photon amplitude involving the electromagnetic form factor of the pion and a gauge-invariant residual amplitude. For the neutral pion, the soft-photon amplitude vanishes. We have analyzed the low-energy behavior of the residual amplitudes in terms of generalized polarizabilities. In this context we have introduced two alternative def\/initions of the generalized polarizabilities, a f\/irst one based on a multipole expansion in the center-of-mass frame, and a second one based on a covariant approach interpreted in the pion Breit frame. The connection between the different approaches has been established. In the framework of ChPT at ${\cal O}(p^4)$, the momentum dependence of the generalized polarizabilities is entirely predicted in terms of the pion mass and the pion-decay constant, i.e., no additional counter-term contribution appears. As in the case of real Compton scattering, the results at ${\cal O}(p^4)$ show a degeneracy of the polarizabilities, $\alpha_L(q^2)=\alpha_T(q^2)=-\beta(q^2)$, which we expect to be lifted at the two-loop level. \section{Acknowledgements} This work was supported by the Deutsche Forschungsgemeinschaft (SFB 443). A.\ L.\ thanks the theory group of the Institut f\"{u}r Kernphysik for the hospitality and support during his stay in Mainz where part of his work was done. \frenchspacing
\section{Introduction} Because M82 is thought to be an archetypical starburst galaxy, many X-ray observations of M82 have been made. The ASCA first observation of M82 was conducted in 1993. Tsuru et al. (1997) analyzed the ASCA spectrum and found that it consists of three components: soft, medium, and hard components. The soft and medium components showing emission lines from various elements are thermal origin at temperatures of $\sim$ 0.3 and $\sim$ 1 keV. The ASCA images of the soft and medium components are extended compared with the ASCA point spread function (PSF). Thus their origin would be the galactic wind driven by stellar winds from massive stars and supernovae. The hard component can be well described by either a power-law model (the photon index is $\Gamma \sim$ 1.7) or a thermal-plasma model (the temperature is $kT \sim$ 14 keV). This component is dominant in the X-ray spectra above the 2 keV band, and its X-ray luminosity in the 2 -- 10 keV band is $\sim\ 3\times10^{40}$ erg/s. Tsuru et al. (1997) compared the ASCA flux in the 2 -- 10 keV band with Ginga (Tsuru 1992) and EXOSAT (Schaaf et al. 1989), and found time variability. Furthermore, the spatial extent of the hard component is consistent with a point source within the ASCA angular resolution. All these may suggest that the origin of the hard component is a low-luminosity AGN of M82. However, since Ginga and EXOSAT are non-imaging detectors, possible contamination from other hard sources is not excluded. Therefore, it is unclear whether the flux change can be attributed to the hard component or not. Since the peak position of the hard component does not agree with the soft component, it is unclear whether the peak of the hard component agrees with the center of M82 where a luminous X-ray point source showing time variability exists (Watson et al. 1984; Collura et al. 1994; Bregman et al. 1995; Strickland et al. 1997). Furthermore, the Ginga spectrum can be fitted with a thermal model but cannot be fitted with a power-law model (Tsuru 1992), which is different from typical X-ray spectra of AGNs. The same conclusion was suggested by Cappi et al. (1998) using BeppoSAX data. Thus the origin of the hard component is still debatable. The first key to reveal the origin of the hard component is to clarify whether it shows time variability or not. The second key is to detect the iron K-line emission and to determine its central energy. The central energy of the line can be direct evidence. For these purposes, we made a monitoring observation of M82 with ASCA in 1996. The total exposure time including the observation in 1993 is about $1.5\times10^5$ s. Throughout this paper, a distance of 3.25 Mpc to M82 is assumed. The number of atoms per hydrogen for the cosmic metal abundance adopted in this paper are $9.77\times10^{-2}$ for He, $3.63\times10^{-4}$ for C, $1.12\times10^{-4}$ for N, $8.51\times10^{-4}$ for O, $1.23\times10^{-4}$ for Ne, $3.80\times10^{-5}$ for Mg, $3.55\times10^{-5}$ for Si, $1.62\times10^{-5}$ for S, $3.63\times10^{-6}$ for Ar, $2.29\times10^{-6}$ for Ca, $4.68\times10^{-5}$ for Fe, and $1.78\times10^{-6}$ for Ni (Anders, Grevesse 1989). \section{Observation and Data Reduction} ASCA observed M82 once in 1993, and observed it nine times in 1996. Though the results of the observation in 1993 have been reported by many authors (Tsuru et al. 1994; Awaki et al. 1996; Tsuru et al. 1996; Moran, Lehnert 1997; Ptak et al. 1997; Tsuru et al. 1997; Dahlem et al. 1998), we reanalyzed the data in 1993 in the same method as for the 1996 data to investigate the time variability of M82 systematically. The dates of these observations are listed in table 1. All of the data were obtained with two solid-state imaging spectrometers (SIS0 and SIS1) and two gas-imaging spectrometers (GIS2 and GIS3) at the foci of four thin-foil X-ray mirrors (XRT) on board the ASCA satellite. Details of the instruments can be found in Burke et al. (1991), Ohashi et al. (1996), Makishima et al. (1996), and Serlemitsos et al. (1995), while Tanaka et al. (1994) gives a general description of ASCA. The SIS data were obtained in the 4-CCD bright mode for the observation in 1993 and in the 1-CCD faint mode for the other observations in 1996. The GIS data were obtained in the normal PH mode. All of the data were screened with the standard selection criteria: data taken in the South Atlantic Anomaly, Earth occultation, and regions of low geomagnetic rigidity are excluded. We also eliminated contamination by the bright Earth, removed hot and flickering pixels from the SIS data, and applied rise-time rejection to exclude particle events from the GIS data. We further applied the ``flare-cut'' criteria to exclude the non X-ray background events as many as possible for the GIS data (Ishisaki et al. 1997). After these screenings, we obtained effective exposure times shown in table 1. \section{Analysis and Results} \subsection{Imaging Analysis} Following Tsuru et al. (1997), we made the SIS images (SIS0 + SIS1) in three energy bands (0.4 -- 0.8 keV, 1.2 -- 1.8 keV, and 3.0 -- 10 keV) for an individual observation. We call them the soft-, medium-, and hard-band images, respectively. The three energy bands represent the three spectral components found by Tsuru et al. (1997). Though the soft-band images show somewhat complex structures, we found that all the medium- and hard-band images are single-peaked and no other significant source exists in the field-of-view. Tsuru et al. (1997) found from the PV data (April 19, 1993) that the radial profiles of the soft- and medium-band images are extended compared with the ASCA PSF, while the hard-band image is consistent with a point source. They also found that the peak position of the hard-band image agrees with that of the medium-band image, but does not agree with that of the soft-band image. We confirmed these results of Tsuru et al. (1997) for all the ASCA data in 1996. For example, we show the SIS images of {\#74049000 (March 22, 1996)} in figure 1. The soft-band peak is at $\sim 0.\hspace{-2pt}'6$ southeast of the hard- and medium-band peak. The 90\% confidence level error circle of the pointing position of ASCA is $\sim 1.\hspace{-2pt}'3$ diameter (Gotthelf 1996). Except for {\#74049090} (November 26, 1996), the peak positions of hard-band images are consistent among observations within $\sim 0.\hspace{-2pt}'8$. The hard-band peak of {\#74049090} is located at $\sim 1.\hspace{-2pt}'4$ southeast of that of the PV data (April 19, 1993), which is still almost at the edge of the error circle. Thus, we conclude that the hard-band peaks were located at the same position within the error, and hence the source at the hard-band peak is identical through all the observations. The position of the hard-band peak determined from the ASCA SIS data is $(\alpha, \delta)_{\rm J2000} = (9^{\rm h}55^{\rm m}52^{\rm s}, 69^\circ40'48'')$. \subsection{Light Curve} From the screened data, we extracted the GIS and SIS light curves for each observation from circular regions centered on M82 in two energy bands (0.7 -- 1.5 keV and 3.0 -- 10.0 keV). The bin width of the light curves is 128 s. The extraction radii are $6'$ for the GIS and $4'$ for the SIS, which are nominal for a bright point source according to ``The ASCA DATA Reduction Guide'' presented by the ASCA Guest Observer Facility\footnote{Please see http://heasarc.gsfc.nasa.gov/docs/asca/abc/abc.html}. Next, we combined the light curves of the SIS0 and SIS1, and the GIS2 and GIS3. We fitted the light curves of the GIS2 + GIS3 with a constant counting rate model. In most cases, the reduced $\chi^2\ (=\chi^2/d.o.f.)$ is less than 1.5, which means no apparent time variability. However, the reduced $\chi^2$ is 2.7 for the GIS light curve of {\#74049030 (April 24, 1996)} in the 3.0 -- 10.0 keV band, though we can see no clear time variability in the 0.7 -- 1.5 keV band at that time (figure 2). The SIS data are consistent with these results of the GIS. As mentioned in the previous section, only one identical source was found in the field-of-view through all the observations. Therefore this result strongly suggests that the hard component of M82 detected by Tsuru et al. (1997) has short-term variability on a time scale of $\sim\ (1 - 2)\times10^{4}$ s, while the soft component has no clear time variability. \subsection{Spectral Analysis} From the screened data, we extracted the SIS and GIS spectra from the same regions as those for the light curves. Then we rebinned the spectra to contain at least 20 counts in each spectral bin to utilize the $\chi^2$ technique. We extracted the background spectra for the GIS from the blank sky data taken during the Large Sky Survey project (e.g. Ueda et al. 1998) with the same data reduction method for the data of M82. We obtained the SIS background spectra for the PV data in 1993 from a source-free region around M82. Because the observations in 1996 were conducted with the 1-CCD faint mode and X-rays from M82 covered the whole region of the CCD chip, we could not define the source-free regions for the observations in 1996. Therefore, to extract the background spectra, we used the data of LSS1988+317, which is a dim source observed in the 1-CCD faint mode in 1996 (Sakano et al. 1998). After applying the same data reduction as that for M82, we extracted the background spectra from a source-free region. The count rates of the spectra in the 0.7 -- 1.5 keV and 3.0 -- 10 keV bands are shown in figure 3. They are background subtracted but not vignetting corrected values. The count rates in the 3 -- 10 keV band show rather large time variability compared with those in the 0.7 -- 1.5 keV band. This suggests that the hard component (Tsuru et al. 1997) has long-term variability on a time scale of $\sim$ a month. Because the detector positions of M82 are almost the same for all the observations in 1996, the vignetting correction does not change the results essentially. \subsubsection{Time variability of the hard component} First, we fitted the SIS0 and SIS1 spectra in the 0.6 -- 10 keV band and the GIS2 and GIS3 spectra in the 0.7 -- 10 keV band simultaneously with the three-temperature thermal plasma model applied in Tsuru et al. (1997). The model can be represented by \begin{equation} N_{\rm H}(\mbox{whole}) \times [\mbox{RS(soft)} + N_{\rm H}(\mbox{medium}) \times \mbox{RS(medium)} + N_{\rm H}(\mbox{hard}) \times \mbox{RS(hard)}], \end{equation} where $N_{\rm H}$ is the equivalent hydrogen column density of an absorbing cold material and RS is the thin thermal plasma model developed by Raymond and Smith (1977) (hereafter RS model). The soft and medium components are assumed to have the same metal abundance, and the abundance ratios among metals of the hard component are fixed to be cosmic. Column densities, temperatures, and metal abundances of the soft and medium components are fixed to the best-fit values of Tsuru et al. (1997); $N_{\rm H}$(whole) = $3.0\times10^{20}$ cm$^{-2}$, $N_{\rm H}$(medium) = 0.0 cm$^{-2}$, $kT$(soft) = 0.32 keV, $kT$(medium) = 0.95 keV, He = C = 1.0 cosmic, N = 0.0 cosmic, O = 0.063 cosmic, Ne = 0.15 cosmic, Mg = 0.25 cosmic, Si = 0.40 cosmic, S = 0.47 cosmic, Ar = Ca = 0.0 cosmic, and Fe = Ni = 0.049 cosmic. Thus the free parameters are the column density, temperature, metal abundance, normalization of the hard component, and the normalizations of the soft and medium components. This 3RS model could fit all the spectra quite well. The best-fit parameters are shown in table 2. The luminosities of the soft and medium components are quite stable, while the hard component has clear time variability. This is shown in figure 4. The hard component could also be fitted with a power-law model (Tsuru et al. 1997). Therefore, we also tried the 2RS + power-law model, which is expressed as \begin{equation} N_{\rm H}(\mbox{whole}) \times [\mbox{RS(soft)} + N_{\rm H}(\mbox{medium}) \times \mbox{RS(medium)} + N_{\rm H}(\mbox{hard}) \times \mbox{Power-law(hard)}]. \end{equation} The soft and medium components are assumed to have the same metal abundances. Column densities, temperatures and metal abundance of the soft and medium components are fixed to the best-fit values of Tsuru et al. (1997); $N_{\rm H}$(whole) = $3.0\times10^{20}$ cm$^{-2}$, $N_{\rm H}$(medium) = $0.0$, $kT$(soft) = 0.31 keV, $kT$(medium) = 0.95 keV, He = C = 1.0 cosmic, N = 0.0 cosmic, O = 0.061 cosmic, Ne = 0.14 cosmic, Mg = 0.25 cosmic, Si = 0.40 cosmic, S = 0.45 cosmic, Ar = Ca = 0.0 cosmic, and Fe = Ni = 0.048 cosmic. Thus the free parameters were the column density, photon index, normalization of the hard component, and the normalization of the soft and medium components. This 2RS + power-law model also could fit all the GIS and SIS spectra. The best-fit parameters are shown in table 3. Only the hard component shows a time variability, which is the same as the results of the 3RS model fitting. Though both the 3RS and 2RS + power-law models could fit the spectra well, we should note that the $\chi^2$ values of the 3RS model are generally smaller than those of the 2RS + power-law model. Figure 5 shows the GIS2 spectra of the highest state (\#74049010; April 15, 1996) and the lowest state (\#74049070; October 14, 1996). This figure also shows that only the hard component has a significant time variability. Therefore, we can obtain the spectrum of the variable component by subtracting the spectrum of the lowest state from the highest state, which is shown in figure 6. The spectrum could be fitted by either the heavily absorbed RS or heavily absorbed power-law model (table 4). \subsubsection{Iron K-line emission} Information concerning the iron K-line emission is one of the essential keys to determine the origin of the hard component. Since the statistics of an individual observation are limited, we added all the GIS spectra. Since the PV data in 1993 obtained in the 4-CCD bright mode in 1993 are thought to be quite different in quality from those obtained in the 1-CCD faint mode in 1996, we added only the data in 1996 for the SIS. Before adding the spectra, we confirmed the accuracy of the gain calibration using the center energy of the silicon K line at 1.86 keV. Then we simultaneously analyzed the added SIS and GIS spectra and the SIS spectra in 1993 above the 4 keV band. First, we fitted the spectra with a thermal bremsstrahlung model. The quality of the fit is good ($\chi^2/d.o.f.$ = 350.4/563), and is shown in figure 7 (a) and table 5. Next, we added a Gaussian line model to the thermal bremsstrahlung (Brems. + 1 Gaussian). The result is given in figures 7 (b), 8 and table 5, and the $\chi^2/d.o.f.$ is 330.8/560. The decrease of the $\chi^2$ value is 19.6, while the decrease of the $d.o.f.$ is 3. Therefore, we can conclude that ASCA detected the iron K line with significant line width of $\sigma \sim 0.3$ keV at a significance level of more than 99.5 \% (Malina et al. 1976). Assuming the line broadening of 0.3 keV results from the outflow motion of the hot gas, the velocity of the gas should be $1.4\times10^9$cm/s, which is much too higher than that estimated (e.g. McKeith et al. 1995). The detected line center energy is between the iron 6.4 keV and 6.7 keV. Thus the obtained broad line suggests a superposition of a few lines, as seen in our galactic center (Koyama et al. 1996). Therefore, we added two narrow Gaussian lines ($\sigma$ = 0) to the bremsstrahlung model (Brems. + 2 Gaussians). The best-fit parameters are shown in table 5. In this case, the two lines can be attributed to the iron 6.4 keV and 6.7 keV lines. Finally, we tried three narrow Gaussian lines plus the bremsstrahlung model in which the center energy is fixed to 6.4 keV, 6.7 keV, and 7.0 keV (Brems. + 3 Gaussians). The results are also shown in table 5. \section{Discussion} \subsection{Identification of the Hard Component} Two bright X-ray sources showing time variability have been detected with the Einstein HRI and ROSAT HRI (Watson et al. 1984; Collura et al. 1994); source 1 and 2 of Collura et al. (1994). We call them source 1 and source 2, hereafter. The position of the hard component determined solely from the ASCA SIS data is $(9^{\rm h}55^{\rm m}52^{\rm s}, 69^\circ40'48'')_{\rm J2000}$ with an error circle of $\sim 40''$ radius in 90\% confidence level (Gotthelf 1996). The error circle includes source 2. However, the position of the hard component does not agree with source 1, whose position is outside the error circle. Next, we determine the precise position of the hard component by comparing the ASCA SIS image with the ROSAT HRI image. Making the ASCA SIS image in the same energy range as the ROSAT HRI and comparing it with the actual ROSAT HRI image is insufficient, because the shape of effective area as a function of X-ray energy is different between the two instruments. Therefore, we make a fake ROSAT HRI image from the ASCA SIS data by taking the energy dependence of the effective area of the two instruments into account, and comparing it with the actual ROSAT HRI image. To fake the HRI image, we first made the four SIS images in the 0.4 -- 0.8 keV, 0.8 -- 1.2 keV, 1.2 -- 1.6 keV, and 1.6 -- 2.0 keV bands. Next, we multiplied each image by the ratio of the effective area of the HRI to that of the SIS in each energy band. For the effective area of the ROSAT HRI, we used the values in Briel et al. (1996). For the SIS image of the 0.4 -- 0.8 keV band, particularly, we multiply it by the ratio of the effective area of the HRI in the 0.1 -- 0.8 keV band to that of the SIS in the 0.4 -- 0.8 keV band, because the SIS has no efficiency below 0.4 keV. Finally, we made the fake HRI image by adding all four images. In figure 9, we show the fake HRI image using the data of {\#74049000 (March 22, 1996)} as an example. From this analysis, we found that the peak position of the fake HRI image agrees with the ASCA hard-band peak. Please note that the fake HRI image almost reflects the ASCA medium-band image because the peak of effective area as a function of X-ray energy of the ROSAT HRI is at $\sim 1.1$ keV (Briel et al. 1996). Because the peak of the fake HRI image should be at the peak of the real HRI image, we conclude the ASCA hard-band peak agrees with the peak of the real HRI image, which is located at the nucleus of M82 (Collura et al. 1994; Strickland et al. 1997). Source 2 is time variable and located at the peak of the HRI image. This source is reported to correspond to a strong 6 cm radio source 41.5+597 (Strickland et al. 1997), which showed a 100 \% drop in flux within a year (Muxlow et al. 1994). We estimate the count rate of the ROSAT HRI of the hard component from the best-fit parameters given in tables 2 and 3 and compare it with source 2. When we apply the RS model for the hard component, the count rates of the ROSAT HRI are estimated to be 0.017 c/s, 0.096 c/s, and 0.0040 c/s for the PV data (April 19, 1993), the highest state (\#74049010; April 15, 1996), and the lowest state (\#74049070; October 14, 1996), respectively. In the case of the power-law model, the count rates are estimated to be 0.012 c/s for the PV data, 0.075 c/s for the highest state, and 0.0040 c/s for the lowest state. The range of these estimated counting rates of the hard component includes the actual ROSAT HRI counting rates of $0.04\sim 0.06$ c/s of source 2. Thus, we conclude that the hard component at the ASCA hard-band peak is identical with source 2. \subsection{Origin of the Hard Component} The hard component has time variability of $10^4$s -- a month with luminosity of $3\times10^{40}$ -- $1\times10^{41}$ erg/s. A collection of discrete sources such as X-ray binaries, a super Eddington source, a young SNR, inverse Compton scattering of IR photons from relativistic electrons which produce the radio emission, hot interstellar medium, and an AGN have been considered so far as an origin of the hard X-ray emission. Since the typical X-ray luminosity of X-ray binaries is $\sim\ 10^{37}$ erg/s, a few hundred binaries are needed to explain the luminosity of the hard X-ray emission of M82. Therefore, a collection of binaries cannot explain the time variability of M82. Since the typical luminosity of super Eddington sources is $\sim\ 10^{40}$ erg/s (Okada et al. 1998), it cannot explain the highest luminosity of M82. A young SNR cannot become as bright as M82 did in our observation. The radio emission is extended as large as $\sim\ 30''$ (Kronberg, Clarke 1978), which corresponds to $\sim\ 1.5\times10^{21}$ cm, while the time variability on a time scale of $\sim\ 1\times10^{4}$ s implies the size of the emitting region to be smaller than $3\times10^{14}$ cm. Therefore the inverse Compton emission cannot explain the time variability. It is also quite an unnatural situation that the hot interstellar medium with temperature of $\sim\ 10$ keV is confined to a region smaller than $3\times10^{14}$ cm. Therefore, an AGN is the only possible origin that can explain the luminosity and time variability. The X-ray luminosity of the hard component is similar to that of low-luminosity AGNs (LLAGN) (Terashima et al. 1998b). We proved that the hard component is located at the center of M82 and spatially unresolved. The residual spectrum obtained by subtracting the spectrum of the lowest state (\#74049070; October 14, 1996) from the highest state (\#74049010; April 15, 1996) suggests the strong absorption, which means that the variable source is embedded in the galactic center region of M82. These also strongly support the LLAGN origin of the hard component. Because the violent starburst activity dominates the AGN activity, there has been no evidence of the AGN in other wavelengths (e.g. Luts et al. 1998), and one can detect the AGN activity only in the hard X-ray band like NGC6240 (Iwasawa, Comastri 1998). Though we could not constrain the center energy of the iron K line, Cappi et al. (1998) determined the center energy to be 6.7 keV, which may suggest the thermal origin of the hard component. However, we should note that some of the LLAGNs have the broad iron K line at 6.7 keV rather than at 6.4 keV (Ishisaki et al. 1996; Terashima et al. 1998a). As mentioned above, the time variability implies the size of the emitting region to be smaller than $3\times10^{14}$ cm. Assuming that the origin of the hard component is the LLAGN and the X-rays are mainly emitted from a region as large as 6 times its Schwarzschild radius, the mass of the central object is estimated to be less than $2\times10^{8}\MO$. The temperature, photon index, and column density showed in tables 2 and 3 are also not constant, though there is no clear correlation between them. However, there is a correlation between the column density and luminosity of the hard component; the column density increases as the luminosity decreases (figure 10). The same tendency can be seen in NGC4151, which is the only known source showing the variable column density (Yaqoob et al. 1993). Though the most plausible origin of the hard component is a LLAGN, the spectral shape is somewhat strange. The $\chi^2$ values of the 3RS model fitting are generally smaller than those of the 2RS + power-law model fitting. The Ginga spectrum could be fitted by the thermal bremsstrahlung model but could not be fitted by the power-law model (Tsuru 1992). Almost the same conclusion was reported by Cappi et al. (1998) using the BeppoSAX data. All these suggest the spectral shape is thermal-like, while typical LLAGN shows power-law spectrum. Therefore, it may also be possible that the origin of the hard component is a new type of accreting X-ray sources. \section{Conclusion} We found time variability of the hard component of M82 detected by Tsuru et al. (1997) at a time scale of $10^{4}$ s -- a month. The luminosity of the hard component in the 0.5 -- 10 keV band ranges from $3\times10^{40}$ erg/s to $1\times10^{41}$ erg/s. The spatial position of the hard component agrees with the luminous X-ray point source at the center of M82 detected with ROSAT and Einstein (Watson et al. 1984; Collura et al. 1994; Bregman et al. 1995; Strickland et al. 1997). The spatial extent of the hard component is consistent with a point source. The residual spectrum obtained by subtracting the spectrum of the lowest state from the highest state suggests the strong absorption feature, which means that the variable source is embedded in the galactic center region of M82. All these strongly suggest that there is a hidden LLAGN in M82, though its spectral shape is different from typical LLAGNs. We also detected a broad iron K line at $6.56^{+0.14}_{-0.14}$ keV or a superposition of a few iron K lines. There is a correlation between the column density and luminosity of the hard component. This first firm evidence of a LLAGN in M82 suggests a link between starbursts and AGN. \par \vspace{1pc}\par The authors thank K. Koyama, S. Ueno, Y. Terashima, and S. Yamauchi for helpful discussion and useful comments. They are also grateful to P. Hilton for careful review of the manuscript. HM is supported by the Special Postdoctoral Researchers Program of RIKEN. The authors also thank the ASCA team members for their support. \vspace{1pc}\par {\it Note added in proof.} -- After the submitting this paper, the authors have become aware of the paper by Ptak and Griffiths (astro-ph/9903372), which obtains the similar results to this paper using the same data independently. \clearpage \section*{References} \re Anders E., Grevesse N. 1989, Geochim. Cosmochim. Acta 53, 197 \re Awaki H, Tsuru T., Koyama K., {\it ASCA} team 1996, in UV and X-Ray Spectroscopy of Astrophysical and Laboratory Plasmas, ed. K. Yamashita, T. Watanabe (Universal Academy Press, Tokyo), p327 \re Bregman J. N., Schulman E., Tomisaka K. 1995, ApJ 439, 155 \re Briel U. G., Aschenbach B., Hasinger G., Hippmann H., Pfeffermann E., Predehl P., Schmitt J. H. M. M., Voges W. et al. 1996, The ROSAT User's Handbook \re Burke B. E., Mountain R. W., Harrison D. C., Bautz M. W., Doty J. P., Ricker G. R., Daniels P. J. 1991, IEEE Trans ED-38, 1069 \re Cappi M., Persic M., Mariani S., Bassabu L., Danese L., Dean A. J., Di Cocco G., Franceschini A. et al. 1998, preprint (astro-ph/9809325) \re Collura A., Reale F., Schulman E., Bregman J. N. 1994, ApJ 420, L63 \re Dahlem M., Weaver K. A., Heckman T. M. 1998, ApJS 118, 401 \re Gotthelf E. 1996, ASCA News No. 4, p31 (ASCA Guest Observer Facility, NASA, Goddard Space Flight Center) \re Ishisaki Y., Makishima K., Iyomoto N., Hayashida K., Kohmura Y., Mushotzky R. F., Petre R., Serlemitsos P. J., Terashima Y. 1996, PASJ, 48, 237, 1996 \re Ishisaki Y., Ueda Y., Kubo H., Ikebe Y., Makishima K., the GIS team 1997, ASCA News No. 5, p26 (ASCA Guest Observer Facility, NASA, Goddard Space Flight Center) \re Iwasawa K., Comastri A. 1998, MNRAS 297, 1219 \re Koyama K., Maeda Y., Sonobe T., Takeshima T., Tanaka Y., Yamauchi S. 1996, PASJ 48, 249 \re Kronberg P. P., and Clarke J. N. 1978, ApJ 224, L51 \re Lutz D., Kunze D., Spoon H. W. W., Thornley M. D. 1998, A\&A 333, L75 \re McKeith C. D., Greve A., Downes D., Prada F. 1995, A\&A 293, 703 \re Makishima K., Tashiro M., Ebisawa K., Ezawa H., Fukazawa Y., Gunji S., Hirayama M., Idesawa E. et al. 1996, PASJ 48, 171 \re Malina R., Lampton M., Bowyer S. 1976, ApJ 209, 678 \re Moran E. C., Lehnert M. D. 1997, ApJ 478, 172 \re Muxlow T. W. B., Pedlar A., Wilkinson P. N., Axon D. J., Sanders E. M., de Bruyn A. G. 1994, MNRAS 266, 455 \re Ohashi T., Ebisawa K., Fukazawa Y., Hiyoshi K., Horii M., Ikebe Y., Ikeda H., Inoue H. et al. 1996, PASJ 48, 157 \re Okada K., Dotani T., Makishima K., Mitsuda K., Mihara T. 1998, PASJ 50, 25 \re Ptak A., Serlemitsos P., Yaqoob T., Mushotzky R., Tsuru T. 1997, AJ 113, 1286 \re Raymond J. C., Smith B. W. 1977, ApJS 35, 419 \re Sakano M., Koyama K., Tsuru T., Awaki H., Ueda Y., Takahashi T., Akiyama M., Ohta K., Yamada T. 1998, ApJ 505, 129 \re Schaaf R., Pietsch W., Biermann P. L., Kronberg P. P., Schmutzler T. 1989, ApJ 336, 722 \re Serlemitsos P. J., Jalota L., Soong Y., Kunieda H., Tawara Y., Tsusaka Y., Suzuki H., Sakima Y. et al. 1995, PASJ 47, 105 \re Strickland D. K., Ponman T. J., Stevens L. R. 1997, A\&A 320, 378 \re Tanaka Y., Inoue H., Holt S. S. 1994, PASJ 46, L37 \re Terashima Y., Kunieda H., Misaka K., Mushotzky R. F., Ptak A. F., Reichert G. A. 1998a, ApJ 503, 212 \re Terashima Y., Kunieda H., Serlemitsos P. J., Ptak A. 1998b, in THE HOT UNIVERSE (IAU symposium No. 188), ed. K. Koyama, S. Kitamoto, M. Itoh (Kluwer Academic Publishers, Dordrecht), p444 \re Tsuru T. 1992, PhD Thesis, The University of Tokyo \re Tsuru T., Hayashi I., Awaki H., Koyama K., Fukazawa Y., Ishisaki Y., Iwasawa K., Ohashi T. et al. 1994, in New Horizon of X-Ray Astronomy, ed. F. Makino, T. Ohashi (Universal Academy Press, Tokyo), p529 \re Tsuru T., Hayashi I., Awaki H., Koyama K., Fukazawa Y., Ishisaki Y., Iwasawa K., Ohashi T. et al. 1996, in X-ray Imaging and Spectroscopy of Cosmic Hot Plasmas, ed. F. Makino, K. Mitsuda (Universal Academy Press, Tokyo), p157 \re Tsuru T., Awaki H., Koyama K., Ptak A. 1997, PASJ 49, 619 \re Ueda Y., Takahashi T., Inoue H., Tsuru T., Sakano M., Ishisaki Y., Ogasaka Y., Makishima K. et al. 1998, Nature 391, 866 \re Watson M. G., Stanger V., Griffiths, R. E. 1984, ApJ 286, 144 \re Yaqoob T., Warwick R. S., Makino F., Otani C., Sokoloski J. L., Bond A., Yamauchi M. 1993, MNRAS 262, 435 \clearpage \centerline{Figure Caption} \bigskip \begin{fv}{1}{} {SIS images of M82 from {\#74049000 (March 22, 1996)} without background subtraction and vignetting correction, (a) in the 0.4 -- 0.8 keV band, (b) in the 1.2 -- 1.8 keV band, and (c) in the 3.0 -- 10 keV band. The pixel size is $6.\hspace{-2pt}''4$, and the images have been smoothed with a Gaussian distribution of $\sigma$ = 2 pixel. The contour levels are 0.1, 0.2, 0.3, ... c/pixel for (a), and 1, 2, 3,... c/pixel for both (b) and (c). Dots, triangles, and stars denote the peaks of the soft-, medium- and hard-band images, respectively. } \end{fv} \begin{fv}{2}{} {Light curves of the GIS2+3 during the observation of \#74049030 (April 24, 1996). Each bin width is 128 s. The horizontal axis is time after the start of the observation. No background subtraction is applied. The background levels (0.0079 c/s for the 0.7 -- 1.5 keV band and 0.012 c/s for the 3.0 -- 10.0 keV band) are shown by the dashed lines.} \end{fv} \begin{fv}{3}{} {Count rates of the GIS2+3 in the 0.7 -- 1.5 keV and 3.0 -- 10 keV bands. The error bars are smaller than the data points.} \end{fv} \begin{fv}{4}{} {Time variability of each component in the 3RS model. Stars, crosses, and circles show the hard, medium, and soft components, respectively.} \end{fv} \begin{fv}{5}{} {GIS2 spectra of the highest state (\#74049010; April 15,1996) and the lowest state (\#74049070; October 14, 1996).} \end{fv} \begin{fv}{6}{} {Residual spectra of the SIS0, SIS1, GIS2, and GIS3 obtained by subtracting the lowest state (\#74049010; April 15, 1996) from the highest state (\#74049070; October 14, 1996). The lines show the best-fit heavily absorbed RS model.} \end{fv} \begin{fv}{7}{} {The composite SIS and GIS spectra and the SIS spectra in 1993 above the 4 keV band. The solid lines show the best-fitting models: (a) thermal bremsstrahlung model, and (b) thermal bremsstrahlung + Gaussian line model. } \end{fv} \begin{fv}{8}{} {Confidence contours at $\Delta\chi^2$ = 2.3, 4.61, and 9.21 for the line center energy and equivalent width of the iron K line.} \end{fv} \begin{fv}{9}{} {The fake ROSAT HRI image made of the SIS images of {\#74049000 (March 22, 1996)} without background subtraction and vignetting correction. The pixel size is $6.\hspace{-2pt}''4$, and the image has been smoothed with a Gaussian distribution of $\sigma$ = 2 pixel. The contour levels are 1, 2, 3, ... c/pixel. The symbols are the same as in Fig. 2. } \end{fv} \begin{fv}{10}{} {Column density of the hard component plotted against the luminosity of the hard component in the 2 - 10 keV band. The column density and luminosity are evaluated with the 3RS model.} \end{fv} \clearpage \begin{table}[t] \begin{center} Table~1.\hspace{4pt}Log of the observations. \end{center} \begin{tabular*}{\textwidth}{@{\hspace{\tabcolsep}\extracolsep{\fill}}p{6pc}ccccccc} \hline \hline Sequence &\multicolumn{2}{c}{Date} & &\multicolumn{4}{c}{Exposure} \\ \cline{2-3} \cline{5-8} &UT &MJD & &SIS0 &SIS1 &GIS2 &GIS3\\ &(dd/mm/yy) & & &(s) &(s) &(s) &(s)\\ \hline PV \dotfill &19/04/1993 &49096.5 & &16254 &16521 &26694 &26646 \\ 74049000 \dotfill &22/03/1996 &50164.5 & &13565 &13565 &12932 &12924 \\ 74049010 \dotfill &15/04/1996 &50188.5 & &6930 &6930 &8668 &8668 \\ 74049020 \dotfill &21/04/1996 &50194.5 & &12395 &12343 &13140 &13138 \\ 74049030 \dotfill &24/04/1996 &50197.5 & &29800 &29902 &33294 &33288 \\ 74049050 \dotfill &13/05/1996 &50216.5 & &8160 &8088 &9688 &9688 \\ 74049060 \dotfill &16/05/1996 &50219.5 & &11545 &11583 &14386 &14390 \\ 74049070 \dotfill &14/10/1996 &50370.5 & &8941 &8941 &7612 &7612 \\ 74049080 \dotfill &14/11/1996 &50401.5 & &12033 &12103 &12498 &12496 \\ 74049090 \dotfill &26/11/1996 &50413.5 & &8762 &8730 &9792 &9760 \\ \hline \end{tabular*} \end{table} \clearpage \begin{table}[t] \begin{center} Table~2.\hspace{4pt}Results of the 3RS model fitting. \end{center} \scriptsize \begin{tabular*}{1.1\textwidth}{@{\hspace{\tabcolsep}\extracolsep{\fill}}p{4pc}ccccccccccc} \hline \hline Sequence &$N_{\rm H}$ &kT &Abundance &\multicolumn{3}{c}{$F_X^*$($10^{-11}$ erg/s/cm$^2$)} & &\multicolumn{3}{c}{$L_X^{\dag}$($10^{40}$ erg/s)} & $\chi^2/d.o.f.$\\ \cline{5-7}\cline{9-11} &$10^{22}$ cm$^{-2}$ &keV &cosmic &soft &medium &hard$^{\S}$ & &soft &medium &hard$^{\S}$\\ \hline PV \dotfill &$2.20^{+0.34}_{-0.33}$ &$11.2^{+2.0}_{-1.6}$ &$0.18^{+0.08}_{-0.06}$ &0.44 &0.70 &2.0 (1.9) & &1.8 &1.6 &4.2 (2.9) &1053.6/1022 \\ 74049000 \dotfill &$1.89^{+0.31}_{-0.29}$ &$4.47^{+0.41}_{-0.38}$ &$0.036^{+0.042}_{-0.036}$ &0.34 &0.88 &2.2 (2.1) & &1.4 &2.0 &5.6 (3.2) &796.2/840 \\ 74049010 \dotfill &$1.35^{+0.25}_{-0.23}$ &$10.2^{+1.5}_{-1.3}$ &$0.026^{+0.071}_{-0.026}$ &0.41 &0.76 &5.5 (5.2) & &1.6 &1.7 &11 (7.4) &859.1/870 \\ 74049020 \dotfill &$1.38^{+0.19}_{-0.18}$ &$8.17^{+0.81}_{-0.73}$ &$0.033^{+0.044}_{-0.033}$ &0.43 &0.73 &4.5 (4.2) & &1.7 &1.7 &9.1 (6.1) &1097.7/1065 \\ 74049030 \dotfill &$1.47^{+0.13}_{-0.13}$ &$7.40^{+0.41}_{-0.38}$ &$0.036^{+0.026}_{-0.027}$ &0.42 &0.74 &4.4 (4.1) & &1.7 &1.7 &9.2 (6.0) &1739.1/1594 \\ 74049050 \dotfill &$1.74^{+0.56}_{-0.54}$ &$16.0^{+7.0}_{-4.6}$ &$0.00^{+0.14}_{\mbox{$\cdot \cdot \cdot$}}$ &0.34 &0.82 &2.3 (2.2) & &1.4 &1.9 &4.4 (3.2) &634.0/596 \\ 74049060 \dotfill &$1.58^{+0.30}_{-0.28}$ &$8.08^{+1.14}_{-0.90}$ &$0.075^{+0.060}_{-0.063}$ &0.34 &0.85 &2.8 (2.6) & &1.4 &1.9 &5.8 (3.9) &929.9/889 \\ 74049070 \dotfill &$3.66^{+0.99}_{-0.91}$ &$5.70^{+2.20}_{-1.32}$ &$0.22^{+0.13}_{-0.12}$ &0.33 &0.82 &1.2 (1.1) & &1.3 &1.9 &3.1 (2.0) &338.1/315 \\ 74049080 \dotfill &$2.84^{+0.73}_{-0.72}$ &$11.7^{+6.8}_{-3.2}$ &$0.12^{+0.13}_{-0.12}$ &0.36 &0.77 &1.7 (1.7) & &1.5 &1.8 &3.7 (2.6) &544.7/479 \\ 74049090 \dotfill &$1.76^{+0.75}_{-0.79}$ &$11.9^{+5.7}_{-3.1}$ &$0.090^{+0.13}_{-0.090}$ &0.37 &0.80 &2.6 (2.5) & &1.5 &1.8 &5.1 (3.6) &490.9/466 \\ \hline \end{tabular*} \noindent All the errors are described at 90\% confidence limits. \noindent $*$ Flux in the 0.5 -- 10 keV band. \noindent $\dag$ Unabsorbed luminosity in the 0.5 -- 10 keV band. \noindent $\S$ Values in parentheses are in the 2 -- 10 keV band. \end{table} \begin{table}[t] \begin{center} Table~3.\hspace{4pt}Results of the 2RS + power-law model fitting. \end{center} \scriptsize \begin{tabular*}{1.05\textwidth}{@{\hspace{\tabcolsep}\extracolsep{\fill}}p{4pc}cccccccccc} \hline \hline Sequence &$N_{\rm H}$ &Photon Index &\multicolumn{3}{c}{$F_X^*$($10^{-11}$ erg/s/cm$^2$)} & &\multicolumn{3}{c}{$L_X^{\dag}$($10^{40}$ erg/s)} &$\chi^2/d.o.f.$\\ \cline{4-6}\cline{8-10} &$10^{22}$ cm$^{-2}$ & &soft &medium &hard$^{\S}$ & &soft &medium &hard$^{\S}$ \\ \hline PV \dotfill &$2.87^{+0.41}_{-0.40}$ &$1.79^{+0.08}_{-0.08}$ &0.42 &0.75 &2.0 (1.9) & &1.7 &1.7 &5.1 (3.1) &1080.4/1023 \\ 74049000 \dotfill &$3.07^{+0.40}_{-0.37}$ &$2.51^{+0.11}_{-0.10}$ &0.30 &0.98 &2.2 (2.1) & &1.3 &2.2 &11 (3.7) &817.2/841 \\ 74049010 \dotfill &$1.90^{+0.31}_{-0.29}$ &$1.83^{+0.08}_{-0.07}$ &0.37 &0.88 &5.5 (5.2) & &1.5 &2.0 &13 (7.9) &876.9/871 \\ 74049020 \dotfill &$2.00^{+0.24}_{-0.23}$ &$1.96^{+0.06}_{-0.06}$ &0.40 &0.84 &4.5 (4.2) & &1.6 &1.9 &12 (6.5) &1121.2/1066 \\ 74049030 \dotfill &$2.25^{+0.17}_{-0.17}$ &$2.04^{+0.04}_{-0.04}$ &0.37 &0.89 &4.4 (4.1) & &1.5 &2.0 &12 (6.5) &1832.8/1595 \\ 74049050 \dotfill &$2.35^{+0.70}_{-0.64}$ &$1.69^{+0.14}_{-0.13}$ &0.32 &0.87 &2.3 (2.2) & &1.3 &2.0 &5.1 (3.3) &637.1/597 \\ 74049060 \dotfill &$2.23^{+0.38}_{-0.34}$ &$1.96^{+0.08}_{-0.09}$ &0.31 &0.92 &2.8 (2.7) & &1.3 &2.1 &7.5 (4.1) &949.6/890 \\ 74049070 \dotfill &$4.27^{+1.17}_{-1.06}$ &$2.10^{+0.26}_{-0.24}$ &0.32 &0.83 &1.2 (1.2) & &1.3 &1.9 &4.3 (2.1) &347.5/316 \\ 74049080 \dotfill &$3.65^{+0.85}_{-0.83}$ &$1.80^{+0.18}_{-0.16}$ &0.34 &0.81 &1.7 (1.7) & &1.4 &1.8 &4.6 (2.8) &549.5/480 \\ 74049090 \dotfill &$2.75^{+0.77}_{-0.79}$ &$1.82^{+0.16}_{-0.15}$ &0.32 &0.90 &2.6 (2.5) & &1.3 &2.1 &6.5 (3.9) &501.2/467 \\ \hline \end{tabular*} \noindent All the errors are described at 90\% confidence limits. \noindent $*$ Flux in the 0.5 -- 10 keV band. \noindent $\dag$ Unabsorbed luminosity in the 0.5 -- 10 keV band. \noindent $\S$ Values in parentheses are in the 2 -- 10 keV band. \end{table} \clearpage \begin{table}[t] \begin{center} Table~4.\hspace{4pt}The results of the fitting of the residual spectra. \end{center} \scriptsize \begin{tabular*}{\textwidth}{@{\hspace{\tabcolsep}\extracolsep{\fill}}p{10cm}cc} \hline \hline &RS &power-law \\ \hline $N_{\rm H}$ ($10^{22}$ cm$^{-2}$) &$1.30^{+0.12}_{-0.10}$ &$1.54^{+0.16}_{-0.14}$\\ Temperature (keV) \dotfill &$11.7^{+2.6}_{-2.1}$ &$\cdot\cdot\cdot$\\ Abundance (cosmic) \dotfill &$0.00^{+0.088}_{\cdot\cdot\cdot}$ &$\cdot\cdot\cdot$\\ Photon Index \dotfill &$\cdot\cdot\cdot$ &$1.72^{+0.09}_{-0.08}$\\ Flux (0.5--10keV) (erg/s/cm$^2$) \dotfill &$4.5\times10^{-11}$ &$4.6\times10^{-11}$\\ Flux (2--10keV) (erg/s/cm$^2$) \dotfill &$4.2\times10^{-11}$ &$4.3\times10^{-11}$\\ Luminosity (0.5--10keV) (erg/s) \dotfill &$8.3\times10^{40}$ &$9.5\times10^{-11}$\\ Luminosity (2--10keV) (erg/s) \dotfill &$5.8\times10^{40}$ &$6.1\times10^{-11}$\\ $\chi^2/d.o.f.$ \dotfill &272.0/274 &279.5/275\\ \hline \end{tabular*} \noindent All the errors are described at 90\% confidence limits. \end{table} \clearpage \begin{table}[t] \scriptsize \begin{center} Table~5.\hspace{4pt}The best-fit parameters of the model fitting to the spectra above the 4 keV band. \end{center} \begin{tabular*}{1.1\textwidth}{@{\hspace{\tabcolsep}\extracolsep{\fill}}p{8pc}ccccccc} \hline\hline &Brems. &Brems. + 1 Gaussian &\multicolumn{2}{c}{Brems.+ 2 Gaussians} &\multicolumn{3}{c}{Brems. + 3 Gaussians}\\ \hline Temperature \dotfill &$12.56^{+1.95}_{-1.33}$ &$10.19^{+1.84}_{-1.05}$ &\multicolumn{2}{c}{$10.71^{+1.43}_{-1.14}$} &\multicolumn{3}{c}{$10.51^{+1.48}_{-1.16}$}\\ Center Energy (keV) \dotfill &\mbox{$\cdot\cdot\cdot$} &$6.56^{+0.14}_{-0.14}$ &$6.34^{+0.12}_{-0.14}$ &$6.68^{+0.17}_{-0.11}$ &$6.4^{\dag}$ &$6.7^{\dag}$ &$7.0^{\dag}$ \\ Line Width ($\sigma$) (keV) \dotfill &\mbox{$\cdot\cdot\cdot$} &$0.30^{+0.16}_{-0.19}$ &$0^{\dag}$ &$0^{\dag}$ &$0^{\dag}$ &$0^{\dag}$ &$0^{\dag}$ \\ Equivalent Width (eV) \dotfill &\mbox{$\cdot\cdot\cdot$} &$121^{+58}_{-61}$ &$39^{+26}_{-25}$ &$52^{+31}_{-30}$ &$45^{+26}_{-25}$ &$39^{+31}_{-31}$ &$22^{+37}_{-22}$\\ $\chi^2/d.o.f.$ \dotfill &350.4/563 &330.8/560 &\multicolumn{2}{c}{330.2/559} &\multicolumn{3}{c}{330.0/560}\\ \hline \end{tabular*} \noindent All the errors are described at 90\% confidence limits. \noindent $\dag$ Fixed. \end{table} \end{document} \clearpage \epsfile{file=fig1a.eps,width=0.8\textwidth} \epsfile{file=fig1b.eps,width=0.8\textwidth} \epsfile{file=fig1c.eps,width=0.8\textwidth} \epsfile{file=fig2.eps,width=0.8\textwidth} \epsfile{file=fig3.eps,width=0.8\textwidth} \epsfile{file=fig4xxx.eps,width=0.8\textwidth} \epsfile{file=fig5.eps,width=0.8\textwidth} \epsfile{file=fig6.eps,width=0.8\textwidth} \epsfile{file=fig7a.eps,width=0.8\textwidth} \epsfile{file=fig7b.eps,width=0.8\textwidth} \epsfile{file=fig8.eps,width=0.8\textwidth} \epsfile{file=fig9.eps,width=0.8\textwidth} \epsfile{file=fig10.eps,width=0.8\textwidth} \end{document}
\section{Introduction} The evolution of a neutron star (NS) in a binary system differs considerably in the case the donor star is a high mass giant star in a HMXB or is a lighter star in a LMXB (see {\it e.g.} Verbunt 1993; Lipunov 1992). It is only in the former case that the formation of a pulsar (slowly rotating at $P>50$ ms) in a relativistic NS-NS binary allows for the accurate estimate of the masses of the two components (Portegies, Zwart \& Yungelson 1998). Collecting all data for the known binary pulsars, Thorsett \& Chakrabarty (1998) found an average mass of $1.35~M_\odot$ (with a very narrow spread $\sigma = 0.04~M_\odot$) for both NSs in these systems. These values are fully compatible with the two hypotheses: that NSs at birth possess a canonical mass of about $1.4~M_\odot$ (Woosley \& Weaver 1986; but see Timmes, Woosley \& Weaver 1996) and that the HMXB phase is shortlived so that at most $10^{-2}~M_\odot$ can be accreted onto the oldest nonpulsating NS (Bhattacharya \& van den Heuvel 1991). On the other hand, it is widely accepted (Phinney \& Kulkarni 1994, van den Heuvel \& Rappaport 1992) that the formation of millisecond pulsars (MSPs; with period $P<10~$ms) takes place inside LMXBs, where an old neutron star is spun--up by accretion torques. The mass load can be significant, in these systems, and, in principle, the NS can reach ultrashort periods $P \la 1$ ms (depending on the equation of state) on a time that is controlled by the evolution of the magnetic field (Possenti {\it et al.} 1998). Thus, there is a close relationship between the spin period $P,$ the gravitational mass and the magnetic field evolution. Cook, Shapiro \& Teukolsky (1994a: CSTa) developed a relativistic code to study the spin evolution of an unmagnetized NS accreting from the inner edge of a Keplerian disk. Here we reproduce their results using a semi--analytical model (described in section $\S~2$). Then we allow for the presence of a magnetosphere around the NS and explore the consequences on the relation between the baryonic mass load $M_{ac}$ and the spin rate $P$, considering two scenarios for the magnetic field evolution. The minimum mass load requested to reach the submillisecond range is computed for a magnetic neutron star in section $\S~3$, adopting different equations of state (EoSs). Observational tests are discussed in section $\S~4$; in $\S~5$ we present our conclusions. \section{Mass loading and spin evolution} In this section we shortly describe the semi--analytical model used to compute the $P~$vs$~M_{ac}$ relation, which is estimated down to limiting periods $\la~1$ millisecond. We first derive a simple equation describing the evolution of the gravitational mass $M_G$ as a function of the baryon load. We then introduce the set of equations for the evolution of the rotational frequency $\Omega$, including rotational effects on the NS radius during recycling. The angular momentum transfer rate is computed including the corrections due to a strong gravitational field. The equations are solved numerically. The code follows the evolution of an NS of initial gravitational mass $M_{G,0} = 1.4 M_{\odot}$ which is accreting at a constant rate $\dot{m}_B$. The initial values of the NS baryonic mass $M_{B,0}$ and of its radius $R_0$ are known once we specify the EoS (see {\it e.g.} Cook, Shapiro \& Teukolsky 1994b; CSTb). In our model, the gravitational and the baryonic masses $M_G$ and $M_B$ of the NS are related, at any given time, by $M_G = M_B (1-\alpha/R),$ where $R$ is the NS radius circumferential radius ({\it i.e.} the proper circumferential length at the NS equator divided by $2\pi$) and $\alpha$ is a constant evaluated using the initial set of parameters. Accordingly, \begin{equation} M_{G} = M_{B} \left[1+\left(\frac{M_{G,0}}{M_{B,0}}-1\right) \left(\frac{R_0}{R}\right)\right] \end{equation} The decrease in radius due to the mass load is described using the simple relation $R \propto M_G^{-1/3}$ (for equilibria of degenerate nonrelativistic neutrons): with this scaling we estimate the non-rotating NS circumferential radius $R$ at any stage of the accretion. The evolution equation for $M_G$ thus reads: \begin{equation} \dot{M}_G~=~\psi ~\dot{m}_B~= ~\left[ 1 + \left( \frac{M_{G,0}}{M_{B,0}} - 1 \right) \left( \frac{M_G}{M_{G,0}} \right)^{1/3} \right] \left[1- \frac{M_B}{3 M_G} \left( \frac{M_{G,0}}{M_{B,0}} - 1 \right) \left( \frac{M_G}{M_{G,0}} \right)^{1/3} \right]^{-1} ~\dot{m}_B. \end{equation} The mass increase is accompanied by the increase in the rotational frequency $\Omega$, and at any time (i.e., at any current value of $M_G$) we need to determine whether $\Omega$ is below or above the mass shedding limit $\Omega_{max}$ that is the limiting angular frequency at which the gravitational pull is balanced by the centrifugal forces. $\Omega_{max}$ is estimated using the classical expression \begin{equation} \Omega_{max} = \Omega_K = \left( \frac{G M_G}{R_{\Omega,max}^3} \right)^{1/2} \end{equation} where $R_{\Omega,max}$ denotes the circumferential radius of the rotating NS which in turn depends upon $\Omega_{max}$. $R_{\Omega,max}$ is estimated fitting the results of CSTb: \begin {equation} R_{\Omega,max}=1.5\,\, R_0 \end{equation} where $R_0$ is the radius of the static configuration at the current value $M_G$. In table 1 we compare equation (3) with the results of CSTb, derived using a relativistic code for 13 different EoSs ($M_G = 1.4 M_\odot$). The agreement is $\le 5\%$. The maximum angular speed occurs when the NS equatorial speed reaches the classical Keplerian limit computed taking $M_G$ as gravitational mass and as radius $R_{\Omega,max}$: the rapid rotation inflates and deforms the NS surface. While the non-spherically-symmetric deformations (oblateness effects) are important in determining the radius of the spinning NS, they are almost irrelevant in changing the gravitational field felt by a test particle at the NS equator with respect to the corresponding spherical configuration. This is probably due to the fact that most of the mass is concentrated well inside the external radius, where the deformation effects due to rotation are almost negligible. This is predicted by the so called Roche Model (see {\it e.g.} Shapiro \& Teukolsky 1983), that, in the hypothesis outlined above, conclude that the maximum expansion of a uniformly rotating star along its equator is a factor of 3/2, irrespective of the EoS adopted. If $\Omega<\Omega_{max}$, spin up can proceed and the circumferential radius $R_{\Omega}$ of the rotating NS further increases. $R_\Omega$ is computed using the relation \begin{equation} \log_{10}(R_\Omega/R_0) = {\frac{1-(1-\Omega/ \Omega_{max})^{1/3}}{6.75}} \end{equation} derived fitting the results of CSTb. In Fig.~1 we report the values of $R_\Omega$ vs $\Omega$ from CSTb, for 5 EoSs ($M_B = 1.4 M_\odot)$ and the relation given by equation (5). In the numerical scheme, $\Omega_{max}$ is firstly determined from (3) using equation (4) for $R_{\Omega,max}$. Then (5) is used to calculate $R_\Omega$. \subsection{Unmagnetized case} During the accretion process the specific angular momentum $l_{in}$ of the accreting matter is entirely transferred to the NS. From the angular momentum conservation $d(I_{\Omega} \Omega)/d t = \dot{m}_B l_{in}$ we obtain \begin{equation} \dot{\Omega} = \frac{\dot{m}_B}{I_{\Omega}}~~ \left( l_{in} - \psi \Omega \frac{\partial I_{\Omega}}{\partial M_G} \right) \left( 1 + \frac{\Omega}{I_{\Omega}} \frac{\partial I_{\Omega}}{\partial R_{\Omega}} \frac{\partial R_{\Omega}}{\partial \Omega} \right) ^{-1} \end{equation} where $I_{\Omega}$ is the moment of inertia of the neutron star. Taking into account the inflation of the radius caused by the rapid rotation, we extrapolate a result of Ravenhall \& Pethick (1994) obtained for static stars: \begin{equation} I_{\Omega} = 0.21 M_G R_{\Omega}^2 \left( 1 - \frac{2 G M_G}{R_{\Omega} c^2}\right) ^{-1} \end{equation} The value $I_\Omega$ is significantly altered at high rotational rates. Hence, its derivatives with respect to $\Omega$ and $M_G$ cannot be neglected in the equation for $\dot \Omega$. The inner rim of the disk at $r_{in}$ is at few NS gravitational radii, where relativistic gravity is important. Therefore the specific angular momentum $l_{in}$ carried by matter differs from the classical newtonian value $\sqrt{G M_G r_{in}}$. Assuming an accretion disk co-rotating with the NS, and the Kerr metric for the approximate description of gravity around a NS (see Shibata \& Sasaki 1998 for a more detailed analysis of the metric), we obtain from Bardeen (1972): \begin{equation} l_{in} = c \frac{r_{h}^{1/2} (r_{in}^2-2 a r_{h}^{1/2} r_{in}^{1/2} + a^2)}{r_{in}^{3/4} (r_{in}^{3/2}-3 r_{h} r_{in}^{1/2} + 2 a r_{h}^{1/2})^{1/2}} \end{equation} where $r_{h} = G M_G/c^2$ and $a = I_{\Omega} \Omega / M_G c$. For an unmagnetized neutron star the inner rim $r_{in}$ of the accretion disk either skims the NS surface or, for very compact NS, is truncated at the radius of the last stable orbit $r_{ms}$. Therefore the radius $r_{in}$ is defined as: \begin{equation} r_{in} = \left\{ \begin{array}{cc} r_{ms} & \mbox{if $r_{ms} > R_{\Omega}$} \\ R_{\Omega} & \mbox{if $r_{ms} \le R_{\Omega}$} \end{array} \right. \end{equation} \noindent In the Kerr metric $r_{ms}$ is given by: \begin{equation} r_{ms} = r_{h} (3+Z_2-[(3-Z_1)(3+Z_1+2 Z_2)]^{1/2}) \end{equation} with \begin{equation} Z_1 = 1+ [1- (a/r_{h})^2]^{1/3} [(1+ a/r_{h})^{1/3}+(1- a/r_{h})^{1/3}] \end{equation} \begin{equation} Z_2 = [3(a/r_{h})^2 + Z_1^2]^{1/2}. \end{equation} Solving for the differential equations (2) and (6), we determine the NS rotational evolution under the hypothesis of stationary accretion. Our results are in agreement with the results of the relativistic code of CSTa, within $8\%$ even in the high spin regime (see {\it e.g.} Fig.~2). The related computational effort is greatly reduced with respect to a fully relativistic numerical approach. \subsection{Magnetized case} For the case of a magnetized NS, there is another characteristic length, i.e., the magnetospheric radius $r_{mag}$, which has to be compared with $r_{ms}$ and $R_\Omega$ when calculating the value of $r_{in}$. The magnetospheric radius is computed as the product of the Alfven radius (see {\it e.g.} Lipunov 1992) times a factor $\phi \sim 1$, estimated as in Burderi {\it et al.} (1998). For high enough magnetic moment $\mu$, $r_{mag}$ exceeds both $R_{\Omega}$ and $r_{ms}$ and the magnetic coupling between the disk and the star determines the extent of angular momentum transfer. The angular momentum balance relation in this case reads as $d(I_{\Omega} \Omega)/d t = g \dot{m}_B l_{in},$ where the torque function $g=g(\Omega)$ accounts for the details of the interaction between the NS magnetosphere and the accretion disk. When $g=0$ the NS is on the so--called ``spinup line'', where it can load mass without modifying its angular momentum $I_{\Omega} \Omega$. To bracket the uncertainties in the determination of the function $g$, we used two different forms of it, following Wang (1996) and Ghosh \& Lamb (1991). In the integration scheme, equation (6) is modified accordingly, to include the magnetic coupling. \section{The minimum mass load as a function of $P$ for a magnetic NS } We compute the $P$ vs $\ml$ relation for the case of an unmagnetized NS (as in CSTa) considering three selected EoSs. As illustrated in Fig.~3, the spin period $P$ shows a steep dependence on the mass loading just above $\sim 3$ ms and small values of $\ml$ ($0.04-0.06 M_{\odot}$) are sufficient to produce a millisecond NS. By contrast, substantial mass load is requested for spinning the NS below the observed minimum period $ P_{min}^{obs} = 1.558$ ms (PSR B1937+21). When approaching the limiting period $P_{lim}$ for mass shedding, the relation $P~$vs$~\ml$ flattens. is $\sim 0.05 M_\odot.$ As already noted by CSTa, all EoSs (labeled as in Arnett \& Bowers 1977) allow a NS to be spun up to $ P_{min}^{obs} $, but a stiffer EoS requires a higher value of $\ml$: $0.25~M_\odot$ for EoS L (stiff) as compared to $0.1~M_\odot$ for EoS A (soft). We have verified that such figures are only slightly affected by assuming a lower initial gravitational mass for the NS (e.g. 1.30 $M_\odot$ - 1.35 $M_\odot).$ In the case of a magnetized NS, the rate at which angular momentum is transferred to the NS is controlled by the interaction of the magnetosphere with the inner boundary of the accretion disk. Thus, the mass load $\ml$ which is necessary to attain a period $P$ depends upon the magnitude of the $B$-field, on its evolution and on the torque function g($\Omega$). We describe different evolutionary pathways, during recycling, introducing two characteristic times: the spinup time $\tau_{up} = P/{\dot{P}}$ and $\tau_{\mu} = \mu/{\dot{\mu}}$, the timescale of decay of the star's magnetic moment $\mu$. As a guideline we consider (i) a two-steps evolution in which a phase of significant $\mu$-decay (with $\tau_{\mu} \ll \tau_{up}$) precedes the phase of spinup at constant $\mu$ (with $\tau_{\mu} \gg \tau_{up}$); and (ii) evolutionary pathways with $\tau_{\mu}=(6/7)~\tau_{up}.$ This condition describes the idealized situation of a NS sliding parallel to the spinup line (corresponding to a fixed accretion rate) without braking along it. In other words, the decay of $\mu$ is tuned just to allow for the NS to approach but never reach the equilibrium period, where the efficiency of the spinup process drops steeply (Wang 1996, Ghosh \& Lamb 1991). In fact, at the equilibrium period, the star would accrete matter without increasing its rotational rate. It is easy to demonstrate that the mass necessary to attain a final period $P$ in case (i) exceeds that of case (ii) by a factor 4/3. For $\tau_{\mu}\gg \tau_{up}$ even larger amounts of mass are needed. Thus, case (ii) provides a {\it lower limit} on the mass load onto the NS, during the recycling process. The relativistic corrections at short rotational periods and the details of the disk-magnetosphere interaction prevent a simple analytical study of the spin evolution. Hence, we explored numerically the NS evolution. Fig.~4 collects a sample of pathways in the plane $P~$vs$~\ml$ obtained for the equation of state FPS (an intermediate EoS), combining various initial conditions (periods in the range $1-100$ sec, $\mu_{ini}$ between $10^{28} - 10^{30}$ G\,cm$^3$, accretion rates in the interval $0.01 - 1.00$ ${\dot{M}}_{Edd}$) and selecting decay histories according to models (i) and (ii). As illustrated in Fig.~4, the values requested to drive a NS to a final period $P$ depend on the magnetic field evolution. We notice however, that only for NSs with $P>~5$ ms at the end of binary evolution the mass load is sensitive to the history of the $\mu$-field. For a magnetized NS the demand of accreted matter can at most halve, relative to the value inferred for a unmagnetized NS. If evolution proceeds to drive $P$ below $5$ ms, the request of $\ml$ depends weakly on the evolution of $\mu$ since the magnetospheric radius shrinks, becoming comparable to the radius of the last marginally stable orbit or to the physical radius of the star. Therefore, the $P~$vs$~\ml$ relation relaxes to the line that characterizes a unmagnetized NS. \section{Discussion} Millisecond pulsars (MSPs) with $P<~1$ ms have not been detected so far. Though such sources are beyond the sensitivity limits of the radio surveys conducted until now, available data may already place constraints on the distribution of MSPs at periods shorter than $ P_{min}^{obs} $. Through consideration of surveys sensitivities and known selection effects, Cordes \& Chernoff (1997) carried out a likelyhood analysis and found that the best-fit models are those increasing towards short periods with a best-fit minimum period that lies only slightly below $ P_{min}^{obs} $. They in addition found a 95\% chance that the fastest MSPs are slower than 1 ms, and a 1\% chance that they are as fast as 0.65 ms. Recently, Possenti {\it et al.} (1998) carried out a population synthesis calculation to determine the fraction of neutron stars with periods shorter than $ P_{min}^{obs} $ relative to those having $P> P_{min}^{obs} $, for $\mu >7.3 \times 10^{25}$Gcm$^{3}$ (the minimum magnetic moment observed so far). They found that the process of recycling in low mass binaries can lead to a distribution of periods extending below $ P_{min}^{obs} $ (for the soft EoS a maximum is found to develop about 1 ms; Possenti {\it et al.} 1999, in preparation). If future observations (as those undergoing at the Northern Cross in Medicina: D'Amico {\it et al.} 1998) will reveal such pulsars, the underlying NSs should have accreted at least $0.3 M_\odot$. Few estimates of the masses of MSPs in low mass binaries are now available and are reported in Fig.~5. The data, taken from Thorsett \& Chakrabarty (1998), are all compatible with the $P~$vs$~\ml$ relation derived in $\S~3$. In particular for PSR B1855+09, the MSP with a mass estimate based on GR alone ($M_G = 1.41 \pm 0.1 M_\odot$), the (minimum) mass requested to reach the observed period ($P=5.362$ ms) would be of at most $0.06 M_\odot$ (for the stiff EoS and an initially low value of $\mu$). Thorsett \& Chakrabarty (1998) found that data obtained collecting the mass measurements of all the known radio pulsars in binary systems are consistent with a remarkably narrow underlying gaussian mass distribution, with $M_{G}=1.35 \pm 0.04 M_\odot$. Their sample of objects contains however five relativistic NS-NS binary systems that contaminates the statistics of MSP binaries, narrowing the value of $M$ and in turn of $\ml$. The progenitors of the NS-NS systems are massive X-ray binaries, a population which does not lead to the formation of fast spinning pulsars. If the proposed value of $\ml~<0.1~M_\odot$ is representative of the MSPs observed sample, this figure is suggestive that a fast decay of the magnetic field has occurred in these systems (not only in PSR B1855+09). This is not in contraddiction with our estimates on $\ml$ because only for those MSPs recycled to limiting periods $P< P_{min}^{obs} $ the demand of mass load would exceed that of Thorsett \& Chakrabarty. Interestingly, a suggestion of the existence of NSs with masses larger than $1.4 M_{\odot}$ arises also from a recent model (Stella \& Vietri 1999) accounting for the pair kHz-QPOs observed in many LMXB sources (see {\it e.g.} van der Klis 1998 for a review). Stella \& Vietri interprete the upper QPO frequency $\nu_2$ as due to matter inhomogeneities orbiting the NS at the inner disk boundary, while the lower QPO frequency $\nu_1$ is produced by the periastron precession at the inner edge of the accretion disk. In this framework, the observed clustering of the differences $\Delta \nu=\nu_2 - \nu_1$ around 250 - 360 Hz is naturally explained provided the gravitational mass of the typical accreting NSs is $\sim 2~M_\odot.$ Moreover the detection of millisecon X-ray pulsations from SAX J1808.4-3658 at different luminosity levels of the source suggests the presence of a massive NS in this system as discussed by Burderi and King (1998). \section{Conclusion} So far, there is no observational evidence against the existence of submillisecond NSs. In the evolutionary scheme adopted, the formation of submillisecond NSs is a natural outcome if a minimum mass of $\simeq 0.25-0.30 M_\odot$ is accreted and if the magnetic moment decays down to values $10^{25}-10^{26}$ Gcm$^3$ {\it before} the end of recycling. Only highly non-conservative scenarios may indeed prevent the transfer of $\ml \ge 0.3~M_\odot$ in low-mass binaries. In fact, evolutionary considerations shows that the mass transfer occurs either because of the nuclear evolution or because angular momentum losses of a $\sim 1~M_\odot$ companion (see {\it e.g.} Burderi, King \& Wynn 1996; Webbink, Rappaport \& Savonije 1983). A more massive companion ($M \ga 2~M_\odot$ -- if we retain the hypothesis that the NS mass at birth is $\sim 1.4~M_\odot$) would determine a dynamically unstable mass transfer probably ending into a common envelope phase (see Kalogera \& Webbink 1996; 1998). On the other hand, a low mass companion ($M \la 1~M_\odot$) would either underfill its Roche lobe, because its nuclear evolution timescale is longer than the age of the Galaxy for initial orbital periods $\ga 0.5$ days, or, for shorter initial orbital periods, would fills its Roche lobe during its main sequence lifetime, resulting in a near--permanent faint LMXB whose compact object will never appear as a radiopulsar. Therefore, in order to produce a fastly rotating NS, we are restricted to a quite narrow range of companion masses, (namely $1-2~M_\odot$), and of initial orbital periods (namely $0.5-10~$ days). For such binary systems, typical timescales for intense accretion are $\la 10^{8}$ years. Once the mass--transfer ceases, the endpoint is often a circular binary containing a millisecond radio pulsar and a low--mass ($\sim 0.2-0.3 M_\odot$) white dwarf companion, the latter being the degenerate helium core of the sub-giant which transferred mass to the compact object (see {\it e.g.} Burderi, King \& Wynn 1998). This implies that the companion has lost $\ga 1~M_\odot$ during the evolution of the system. Even taking into account the case of an highly non-conservative mass transfer phase, it is difficult to believe that the NS in these systems has accreted less than $0.3~M_\odot$. There might be a number of physical processes that inhibit the formation of sub-MSPs during binary evolution. (a) Slow decay of the magnetic moment would relent the spinup process thus requiring a amount of mass exceedingly large. In addition, precocious freezing of $\mu$ (above $10^{27}$ G cm$^{3}$; as discussed in Konar \& Bhattacharya 1997; 1998) may prevent the formation of very fastly spinning NSs. (b) Long term variation of the accretion rate may be important. The terminal phases of LMXB evolution are still uncertain depending on the Roche-Lobe feeding mechanisms. If the accretion rate ceases on a time comparable to the evolutionary time scale, the propeller effect may cause an increase of the spin period above $ P_{min}^{obs} $. Approximately, the ratio of the spin down time $\tau_{sd}$ by propeller to $\tau_{up}$ scales as \begin{equation} \frac{\tau_{sd}}{\tau_{up}}~= ~\left(\frac{{\dot{m}}_{up}}{{\dot{m}}_{sd}}\right)^{5/7}~ ~\left(\frac{\mu_{up}}{\mu_{sd}}\right)^{2/7}~ ~\frac{1}{\xi} \end{equation} where indexes $up$ and $sd$ refer to spinup and spindown, and $\xi$ is a parameter measuring the efficiency of the extraction of angular momentum from the NS during the propeller phase ($\xi=1.0$ if the efficiency of spin up and down are equal). For spindown to be negligible, that ratio needs to exceed unity, requiring a rather rapid turn off of the mass transfer rate (${\dot{m}}_{sd} \ll {\dot{m}}_{up}$) (Possenti {\it et al.} in preparation). (c) Similar considerations holds in case in which the binary is subject to disk instabilities (Cannizzo 1993 and references therein), leading to long periods of low level of activity -- referred in literature as quiescence -- interrupted by brief periods of outbursts (van Paradijs 1996, King 1998). For the known soft X-ray transients (SXTs: see Campana {\it et al.} 1998 for a complete review) the first factor in equation (13) is $\sim 100$, whilst the ratio of $\mu$'s stays nearly constant during a cycle. If the quiescence phase lasts $\sim 100/\xi$ times longer than the outburst phase, spin down would prevail. The off-on time ratio of observed recurrent transients varies from source to source and is about $10-100.$ It is thus difficult to assess the relevance of this process in affecting the spin evolution and the statistical properties of the MSP population. Recently Li, van den Heuvel \& Wang (1998) argued that unsteady mass transfer in PSR J1455-3330 may have occurred, and explored evolutionary tracks allowing for spindown by propeller during the long quiescent intervals that the binary would experience if fed through a thermally unstable keplerian disk. The discovery of SAX J1808-3658 might challenge this picture. (d) For NSs with hot interiors, the viscosity of the nuclear matter may become inefficient in damping the recently discovered $r-$mode instabilities of the star (Andersson 1998). The related emission of gravitational waves (Stergioulas 1998) could limit the spin of a NS to a small fraction ($\sim 0.2 - 0.3$) of its mass shedding rotational rate. For young NSs with $T > 10^9$, this effect could spindown the star at $P>10$ ms in a very short time ($\sim 1$ yr) (see {\it e.g.} Lindblom, Owen \& Morsink 1998). Andersson, Kokkotas \& Stergioulas (1998) claimed that a similar mechanism could occur during the LMXB-phase, thus preventing the spinup of a NS below $\sim P_{min}^{obs} $. However, the difficulties in the calculations of the values of the shear and bulk viscosity for superfluid nuclear matter at $T \sim 10^7-10^{8.5}$ K (relevant for the accretion phase) leave a large uncertainty in the role of this instability in the recycling process. In addition, crustal deformations induced by accretion can excite emission of gravitational waves inhibiting the spin-up process to period shorter than $\sim 3$ ms (Bildsten 1998).
\section{Introduction} Flux density limited radio source samples are known to contain a large portion of compact sources with high frequency spectral index $\alpha$ $\ge$ 0.5 (S$_{\nu} \sim \nu^{-\alpha}$) and angular sizes below 2$''$. The corresponding projected linear sizes are typically $\le$ 15~kpc\footnote{We assume H$_0$ = 100 km~s$^{-1}$~Mpc$^{-1}$ and q$_0=0.5$}. Fanti et al. (1990) have shown that the majority of these CSS sources cannot be larger sources foreshortened by projection effects, which means that their radio emission originates on sub-galactic scales. It is obvious that one should investigate the connection between CSS and larger radio sources. Two scenarios have been proposed that would naturally explain the observed small sizes. First, they could reflect an early stage in the evolution of radio sources. This is the {\it youth scenario} (Phillips $\&$ Mutel 1982; Carvahlo 1985). The second possibility is that the unusual conditions in the interstellar medium of their host galaxies, such as a higher density and/or the presence of turbulence, inhibit the radio source from growing to larger sizes. This is the {\it frustration scenario} (van Breugel et al. 1984). Most CSS sources exhibit double-lobed structures such as seen in classical radio galaxies. This feature is common to both quasars and galaxies. Quite a few of these are symmetric, which gave rise to the terms CSO (compact symmetric objects, with sizes $\le$ 0.5~kpc) and MSO (medium-size symmetric objects, with sizes $>$ 0.5~kpc). These sources are considered as scaled-down versions of larger-sized double radio sources (Fanti et al. 1995). A minority of CSS sources is made up by sources with complex or highly asymmetric structures. These are mostly quasars, with most of their luminosity provided by the jet. The distortions may be caused by jet bending, which is further amplified by strong projection. \par \noindent A plausible organization scheme for these different morphological classes is to identify CSO and MSO with progenitors of large doubles (`baby radio galaxies'), while the asymmetric CSS sources could represent frustrated radio sources. \section{Sample selection} The sample we choose for this study consists of: \par \noindent 1) the sample selected from the 3CR and the Peacock \& Wall (1981; PW) catalogues (Fanti et al. 1995), and \par \noindent 2) nine sources extracted from the B3\,VLA sample (Vi\-gotti et al. 1989). 1) The 3CR/PW sample (38 sources) was established on the basis of the well-known criteria: linear size $<$ 15 kpc, spectral index $\alpha >$ 0.5 and a limit in radio power P $\geq 10^{26.8}$ W/Hz at 178 MHz for the 3CR, or P $\geq 10^{26}$ W/Hz at 2.7 GHz for the PW. High resolution structure information is available for these sources from the literature. In the framework of this project we observed at 230~GHz the 17 sources for which the flux densities at this frequency were not available from literature. 2) The second complete sample of CSS sources was selected from the ``strong'' section of the B3\,VLA catalogue ($S_{408 {\rm MHz}}\geq$ 0.8 Jy) according to the following definitions: $\alpha_{408{\rm MHz}}^{1.4{\rm GHz}} > 0.5$ up to 10~GHz, and a size of $\la$ 15~kpc. Sources with $\alpha_{408{\rm MHz}}^{1.4{\rm GHz}}< 0.5$ have also been included if their spectrum is steeper than $\alpha=0.5$ at $\nu>1.4$~GHz. These are Giga-Hertz-peaked spectrum (GPS) candidate sources. This sample contains 83 objects. The sample has been recently observed with the VLA at 5 and 8.4 GHz, with resolutions of $\sim 0\farcs2$. Additional observations with VLBI arrays are planned. Only about 60\% of these sources have at the moment either a photometric or a spectroscopic redshift, therefore we limited the list to the objects which could be observed at 230~GHz to the latter. Due to scheduling constraints we only observed 9 sources of the B3\,VLA sample. \begin{figure*} \begin{minipage}[h]{8cm} \epsfxsize=8cm \epsfbox [45 190 540 720] {8461.f1a} \end{minipage} \hfill \begin{minipage}[h]{8cm} \epsfxsize=8cm \epsfbox [80 190 575 720] {8461.f1b} \end{minipage} \caption[]{Selected source spectra (left: type a, right: type b; see Sect. 4.3). The $\chi^2_{\rm red}$-values refer to the pure synchrotron loss models. The self-absorbed part of the spectrum has been modelled in a subsequent fit procedure.} \end{figure*} \section{Observations and data reduction} The observations were done with the IRAM~30-m telescope near Granada, Spain between 10th Feb. and 15th Feb., 1998. The receiver we used was the 37-channel bolometer described by Kreysa et al. (1998), with feeds arranged in a hexagon around the central feed, and beam separations of 20$\arcsec$. The sensitivity of each channel was 70\,mJy/s. The beam size was $10\farcs5$ as derived from pointing scans. All our sources were point-like to this beam. Opacity measurements were made via skydips, from which we derived zenith opacities of 0.1--0.8. Despite the high atmospheric attenuation of the signal, the observations (generally done at high source elevations) could be performed satisfactorily, owing to the relatively quiet atmosphere present during our observing run (resulting in a relatively low sky noise). The calibration factor to convert the observed counts into flux densities was determined by observing the source K3--50A, which has a total flux density of 10\,Jy and a peak flux density of 7.6\,Jy for the beam size of our measurements. The measures were finally corrected for atmospheric attenuation and the elevation-dependent gain of the telescope. The pointing was frequently checked by cross-scanning the sources 0923+392, 1144+402, 1308+326, and 1418+546. The pointing accuracy was found to be $\sim\,3\arcsec$. The same sources were used to adjust the focus at regular intervals. The sources were observed in an ON-OFF mode in which the wobbling secondary mirror of the telescope moved at a frequency of 2\,Hz between the source position and positions located at $\pm~60\arcsec$ in azimuth. Each source was observed between 3 and 5 times, with 10 ON-OFF pairs each. The differences of the flux densities between the main and the reference beams were averaged by weighting inversely proportional to the square of the r.m.s. noise of each measurement. The resulting integration time spent on each source was thus typically 50 minutes, and the final r.m.s. noise of each measurement was between 2 and 5\,mJy. In addition, the calibration may be uncertain by 15\% owing to day-to-day variations in the system gain. The data were recorded with all receivers, with the central one pointing at the source. As the angular extents of the target sources are in the range of a few arcseconds, the outer channels contain only sky emission. Hence, they were used to subtract a mean sky level from the central channel. Using this procedure, we assumed that the scale size of the fluctuations of the sky emission is large compared to the total coverage of the bolometer array ($60\arcsec$), which corresponds to 1.5\,m at a distance of 5\,km. Together with the wobbling rate, this ensures efficient rejection of sky noise, since the time scale of atmospheric fluctuations is much larger (see e.g. Altenhoff et al. 1987). In Tab.~1 we list the measured flux densities and 1-$\sigma$ errors, inclusive of noise and calibration uncertainty, at 230~GHz. \begin{table} \caption{Flux densities at 230 GHz} \centering \begin{tabular}{|lccr|}\hline Name & \multicolumn{2}{c}{Pointing centre (B1950)} & \multicolumn{1}{c|}{S$_{230}$}\\ & \multicolumn{1}{c}{$\alpha$}& \multicolumn{1}{c}{$\delta$} & \multicolumn{1}{c|}{[mJy]} \\ \hline 3C-PW & & & \\ 0404+76 & 04 04 00.13 & 76 48 52.50 & 115 $\pm$ 18 \\ 0428+20 & 04 28 06.86 & 20 31 09.13 & 62 $\pm$ 11 \\ 0740+38 3C186 & 07 40 56.77 & 38 00 30.97 & 5 $\pm$ \ph14 \\ 0758+14 3C190 & 07 58 45.05 & 14 23 04.40 & 8 $\pm$ \ph13 \\ 1005+07 3C237 & 10 05 22.04 & 07 44 58.56 & 14 $\pm$ \ph15 \\ 1019+22 3C241 & 10 19 09.38 & 22 14 39.63 & 2 $\pm$ \ph13 \\ 1153+31 & 11 53 44.08 & 31 44 16.00 & 9 $\pm$ \ph13 \\ 1203+64 3C268.3 & 12 03 54.08 & 64 30 18.40 & 1 $\pm$ \ph12 \\ 1225+36 & 12 25 30.77 & 36 51 47.00 & 0 $\pm$ \ph13 \\ 1358+62 & 13 58 58.36 & 62 25 06.71 & 29 $\pm$ \ph16 \\ 1416+06 3C298 & 14 16 38.80 & 06 42 21.30 & 8 $\pm$ \ph14 \\ 1447+77 3C305.1 & 14 47 49.35 & 77 08 46.65 & 2 $\pm$ \ph13 \\ 1517+20 3C318 & 15 17 50.63 & 20 26 52.95 & 5 $\pm$ \ph13 \\ 1634+62 3C343 & 16 34 01.06 & 62 51 41.80 & 12 $\pm$ \ph14 \\ 1637+62 3C343.1 & 16 37 55.29 & 62 40 34.24 & 7 $\pm$ \ph14 \\ 1819+39 & 18 19 42.33 & 39 41 15.00 & 5 $\pm$ \ph14 \\ 1829+29 & 18 29 17.93 & 29 04 58.29 & 9 $\pm$ \ph14 \\ \hline B3\,VLA & & & \\ 0809+404 & 08 09 31.62 & 40 28 02.76 & 11 $\pm$ \ph13 \\ 0810+460B & 08 10 58.61 & 46 05 48.12 & 2 $\pm$ \ph12 \\ 1025+390B & 10 25 49.34 & 38 59 57.55 & 46 $\pm$ \ph18 \\ 1128+455 & 11 28 56.40 & 45 31 24.57 & 4 $\pm$ \ph13 \\ 1159+395 & 11 59 16.35 & 39 35 52.89 & 5 $\pm$ \ph13 \\ 1225+442 & 12 25 15.70 & 44 17 17.23 & 1 $\pm$ \ph13 \\ 1233+418 & 12 33 10.88 & 41 53 38.00 & 11 $\pm$ \ph14 \\ 1242+410 & 12 42 26.39 & 41 04 29.97 & 83 $\pm$ 13 \\ 1350+432 & 13 50 24.04 & 43 14 09.40 & 4 $\pm$ \ph13 \\ \hline \end{tabular} \end{table} \section{Results and discussion} \label{sec:results} For each source we have compiled flux densities at different frequencies from the literature, mostly from K\"uhr et al. (1979) and from the CATS database (Verkhodanov et al. 1997). Our new measures at 230~GHz have been added to the compilation. All flux densities have been brought to the BGPW scale (Baars et al. 1977). Examples of source spectra are shown in Fig.~1. For the fit algorithm described in Sect. 4.2 we have assumed flux densities less than 3$\sigma$ to be upper limits only. Most of the sources show significant departure from the classical power law which describes a zero age transparent synchrotron spectrum from a relativistic electron population with power law energy distribution. The deviations from the power law are of the following type: a) a low-frequency turnover (the most conspicuous deviation); b) a steepening at high frequencies. High-frequency flattening, if any, is quite rare. The above deviations are interpreted as due to synchrotron self-absorption and to particle energy losses, respectively. In order to describe them, we have fitted the compiled flux densities with a synchrotron aged spectrum ${\rm S}_{\rm aged}(\nu)$ (described in the next section), modified by low-frequency absorption, as follows (Pacholczyk 1970): \begin{equation} {\rm S}(\nu)\propto(\nu/\nu_1)^{\alpha+\beta}(1-e^{-(\nu/\nu_1)^{-(\alpha+\beta)}}) \cdot {\rm S}_{\rm aged}(\nu) \label{a} \end{equation} where $\nu_1$ is the frequency at which the optical depth is equal to 1. In case of an homogeneous synchrotron self-absorbed source $\beta = 2.5$, while $\alpha$ is the not aged spectral index in the transparent frequency range. \subsection{The synchrotron aged spectrum model} We assume that the radio source evolution is described by a {\it continuous injection model}, where the sources are continuously replenished by a constant flow of fresh relativistic particles with a power law energy distribution, with exponent $\delta$. It is well known that, under these assumptions, the radio spectrum has a standard shape (Kardashev 1962), with spectral index $\alpha_{\rm inj} = (\delta -1)/2$ below a critical frequency $\nu_{\rm br}$ and $\alpha_{\rm h} = \alpha_{\rm inj} + 0.5$ above $\nu_{\rm br}$. If there is no expansion and the magnetic field is constant, the frequency $\nu_{\rm br}$ (in GHz) depends on the elapsed time since the source formation, $\tau_{\rm syn}$ (in Myrs), the intensity of the magnetic field $B$ (in $\mu$G) and the magnetic field equivalent to the microwave background $B_{\rm CMB} = 3.25 (1+z)^2$ (in $\mu$G) as: \begin{equation} \tau_{\rm syn} = 1610 {\frac{B^{0.5}}{B^2 + B^2_{\rm CMB}}} {\frac{1}{[\nu_{\rm br}(1+z)]^{1/2}}} \label{b} \end{equation} The whole spectral shape cannot be described by an analytic equation, the two behaviours described being only the asymptotical ones, and has to be computed numerically. This model is referred as continuous injection (CI). Fitting the spectral data to the numerically computed CI spectrum, one obtains the break frequency $\nu_{\rm br}$, from which the source age is obtained if the magnetic field is known. \par \noindent This simple model does not consider expansion effects, which may be important if the source is young. So, the simplest variant of the original model is the one in which the radiating particles loose energy through expansion and the magnetic field changes according to flux conservation (Kardashev 1962, CIE). An alternative possibility to be considered, in an expanding source, is that the magnetic field changes less rapidly than in the flux conserving assumption because of a continuous magnetic flux input associated to the fresh particles injection. We set $B \propto t^{-{\rm m}}$, where for m = 2 we get the flux conserving expansion. This is referred as CIm model. In our special case we assume m = 1, consistent with the models applied by Baldwin (1982) or Begelman (1996). Although the theoretical background to these models is in the Kardashev paper, we have decided to present in the Appendix the detailed development. The spectral shapes of the CIE and CIm models have been computed again numerically. The break frequency for the CIE model is sixteen times higher than in the CI case, for equal elapsed time and final magnetic field intensity (Kardashev 1962). In the CIm model, instead, one finds that the break frequency is $(2.5{\rm m}-1)^2$ larger than in the CI model. The asymptotic behaviours at frequencies lower and larger than the break frequency are the same as in the CI model. However the steepening occurs over a broader frequency interval. In order to better emphasize the differences between these three models it is useful to consider, together with the usual flux-frequency representation, the point-to-point spectral index defined as \begin{equation} \alpha(\nu/\nu_{\rm br})=-\frac{d\log {\rm S}_{\rm aged}(\nu/\nu_{\rm br})}{d\log \nu/\nu_{\rm br}} \end{equation} Both representations are shown in Fig.~2. While in the flux-frequency plane the differences are hardly visible, they can be much better traced in the point-to-point spectral index behaviour. Note that the displayed figures only show the theoretical differences. In practice the observed spectra permit only fits in the flux-frequency plane. \begin{figure}[t] \epsfxsize=8cm \epsfbox [45 150 540 720] {8461.f2} \caption[]{The three continuous injection models described in the text. The solid line corresponds to the CI, the dashed line to the CIE and the dot-dashed line to the CIm. On the top the flux density (arbitrary units) is plotted as a function of the ratio $\nu/\nu_{\rm br}$ (flux-frequency plane), on the bottom the point-to-point spectral index is shown as a function of $\nu/\nu_{\rm br}$. The injection spectral index is the same for all three models $\alpha_{\rm inj}=0.625$.} \end{figure} \subsection{The spectral fits} Spectral fits to the spectra have been made with the CI, CIE, and CIm models. They allow us to determine the non-aged spectral index $\alpha_{\rm inj}$ and the break frequency $\nu_{\rm br}$, that, together with the normalization, are the three free parameters characterizing all the models. The best fits are, surprisingly, obtained with the CI model. The other models have steepenings which are too gradual for the majority of the spectra. An example is given in Fig.~3. The reduced $\chi^2_{\rm red}$ of the models including adiabatic expansion and a variable magnetic field is always greater (typically twice) than that of the CI model (Fig.~4). The CI fits appear quite good, even in cases of high $\chi^2_{\rm red}$ values, which, at visual inspection, appear more due to an under-estimation of the flux density errors than to a poor fit of the spectral model on the data. \begin{figure}[t] \epsfxsize=8cm\epsfbox [45 150 540 720] {8461.f3} \caption[]{Three different fit results for the source 2252+12. The solid line corresponds to a pure CI model, the dashed line represents the CIE model, the dot-dashed line stands for the CIm model. Note the different qualities of the fits, expressed by the various $\chi^2_{\rm red}$ values!} \end{figure} The majority of the spectra show a clear break frequency, with a change of slope $\Delta \alpha \sim 0.5$. We stress that there is no evidence for spectral steepening with $\Delta \alpha$ significantly larger than 0.5. Only a few sources are fitted by simple power laws. In these cases $\nu_{\rm br}$ could be either very high ($\ga 100$ GHz) or very low ($\la {\rm few}~100\;{\rm MHz}$). In these sources we have preferred the low frequency choice, since for the high frequency one would have implied abnormally high values for $\alpha_{\rm inj}$ as compared to the other sources. We also note that only a very small fraction of the CSS sources, if any, shows some evidence of flux density excess at high frequency, as it would be caused by a flat spectrum core or by thermal dust emission. Perhaps the only case is 3C138, where the core, known from VLBI observations, shows up in the integrated spectrum at 230 GHz only. It may appear surprising that the fits with the assumed spectral model are so good. In fact the sources of the sample consist of several components, as lobes and jets and hot spots, where physical conditions can differ from one another and therefore also break frequencies may be different. It is likely that the spectrum is dominated by the brighter component(s). In addition, one could imagine that some confusion might have occurred between genuine spectral steepening due to energy losses and the low-frequency turnover due to absorption processes. We feel that this is a minor problem, but it is difficult to quantify it (see, however, next section). The break frequencies range from a few hundred MHz to tens of GHz. At low frequencies, as said above, the limit is set by confusion with the effects of absorption processes. The injection spectral index $\alpha_{\rm inj}$ ranges from 0.35 to 0.8, with $\langle \alpha_{\rm inj} \rangle = 0.63$. The typical errors of the break frequencies and the injection spectral indices as given by the fit algorithm are up to 40\% and 0.05, respectively. The results of the CI fits are compiled in Tab.~2. \begin{table*} \caption{ Physical Parameters } \centering \begin{tabular}{|llrrcrrrc|}\hline \multicolumn{1}{|c}{Source}&\multicolumn{1}{c}{redshift}& \multicolumn{1}{c}{LS} & \multicolumn{1}{c}{$\nu_{\rm br}$}& \multicolumn{1}{c}{$\alpha_{\rm inj}$} & \multicolumn{1}{c}{$B_{\rm eq}$} & \multicolumn{1}{c}{$\tau_{\rm syn}$} &\multicolumn{1}{c}{$ v_{\rm exp}$/c} & \multicolumn{1}{c|}{type}\\ & & \multicolumn{1}{c}{[kpc]}& \multicolumn{1}{c}{[GHz]} & & \multicolumn{1}{c}{[$10^{-3}$G]}& \multicolumn{1}{c}{[$10^{3}$ yrs]} & & \\ \hline 3C-PW & & & & & & & & \\ 0127+23 (3C43) & 1.46 & 9.40 & $>$100\phantom{.0} & 0.76 & 5.0 & 0.3 & 55.0\phantom{1} & b \\ 0134+32 (3C48) & 0.37 & 2.10 & 1.8 & 0.50 & 2.0 & 11.3 & 0.60 & b \\ 0138+13 (3C49) & 0.62 & 3.60 & 2.6 & 0.57 & 7.0 & 1.3 & 4.39 & b \\ 0221+67 (3C67) & 0.31 & 6.80 & 2.2 & 0.56 & 0.7 & 51.0 & 0.22 & a \\ 0223+34 & 1.00 & 3.80 & 24.1 & 0.38 & 9.4 & 0.3 & 24.20 & b \\ 0316+16 & 1.00 & 1.20 & 12.0 & 0.81 & 9.8 & 0.3 & 5.74 & b \\ 0345+33 (3C93.1)& 0.24 & 1.20 & 1.6 & 0.58 & 1.0 & 35.8 & 0.05 & ? \\ 0404+76 & 0.60 & 0.53 & 10.7 & 0.38 & 4.9 & 1.1 & 0.76 & a \\ 0428+20 & 0.22 & 0.45 & 7.2 & 0.38 & 5.8 & 1.2 & 0.59 & b \\ 0429+41 (3C119) & 1.02 & 0.75 & 9.5 & 0.49 & 8.0 & 0.5 & 2.36 & b \\ 0518+16 (3C138) & 0.76 & 2.90 & 5.3 & 0.47 & 1.0 & 16.7 & 0.28 & b \\ 0538+49 (3C147) & 0.55 & 2.40 & 1.6 & 0.44 & 2.5 & 8.3 & 0.47 & b \\ 0740+38 (3C186) & 1.06 & 8.20 & 0.3 & 0.75 & 0.7 & 113.2 & 0.12 & a \\ 0758+14 (3C190) & 1.20 & 14.10 & 30.3 & 0.79 & 0.7 & 10.6 & 2.15 & b \\ 1005+07 (3C237) & 0.88 & 4.50 & 2.4 & 0.59 & 1.2 & 18.1 & 0.40 & a \\ 1019+22 (3C241) & 1.62 & 2.80 & 1.1 & 0.77 & 2.2 & 9.3 & 0.49 & a \\ 1153+31 & 1.56 & 3.20 & 6.9 & 0.69 & 1.7 & 5.4 & 0.95 & a \\ 1203+64 (3C268.3)& 0.37 & 3.90 & 4.4 & 0.66 & 0.7 & 35.4 & 0.18 & a \\ 1225+36 & 1.98 & 0.17 & 3.9 & 0.64 & 18.0 & 0.2 & 1.41 & b \\ 1250+56 (3C277.1)& 0.32 & 4.40 & 0.7 & 0.39 & 0.4 & 204.3 & 0.04 & a \\ 1323+32 & 0.37 & 0.18 & 6.9 & 0.38 & 5.1 & 1.4 & 0.20 & a \\ 1328+30 (3C286) & 0.85 & 14.20 & 7.3 & 0.38 & 13.0 & 0.3 & 77.82 & b \\ 1328+25 (3C287) & 1.06 & 0.40 & 10.7 & 0.47 & 3.5 & 1.7 & 0.39 & b \\ 1358+62 & 0.43 & 0.16 & 71.9 & 0.70 & 9.0 & 0.2 & 1.39 & b \\ 1416+06 (3C298) & 1.44 & 9.10 & $<$0.1 & 0.50 & 1.6 & $>$50\phantom{.0} &$<$0.3\phantom{1} & a \\ 1443+77 (3C303.1)& 0.27 & 5.00 & 0.8 & 0.64 & 0.6 & 110.2 & 0.07 & a \\ 1447+77 (3C305.1)& 1.13 & 9.00 & 0.8 & 0.62 & 0.4 & 151.3 & 0.10 & a \\ 1458+71 (3C309.1)& 0.90 & 7.80 &109.4 & 0.65 & 7.0 & 0.6 & 21.0\phantom{1} & b \\ 1517+20 (3C318) & 0.75 & 7.80 & 4.8 & 0.69 & 1.5 & 9.6 & 1.32 & b \\ 1607+26 & 0.47 & 0.18 & 6.3 & 0.71 & 10.0 & 0.5 & 0.55 & a \\ 1634+62 (3C343) & 0.99 & 0.70 & 3.8 & 0.65 & 2.0 & 6.5 & 0.17 & ? \\ 1637+62 (3C343.1)& 0.75 & 1.30 & 1.7 & 0.62 & 1.5 & 16.0 & 0.13 & a \\ 1819+39 & 0.40 & 3.10 & 3.3 & 0.73 & 6.0 & 1.6 & 3.13 & b \\ 1829+29 & 0.60 & 9.30 & 12.1 & 0.63 & 6.0 & 0.8 & 19.12 & b \\ 2248+71 (3C454.1)& 1.84 & 6.60 & 0.4 & 0.69 & 1.5 & $\ga$26.5& $<$0.40 & a \\ 2249+18 (3C454) & 1.76 & 2.10 & 11.9 & 0.72 & 5.0 & 0.8 & 4.29 & b \\ 2252+12 (3C455) & 0.54 & 13.70 & 1.7 & 0.62 & 0.2 & 348.2 & 0.06 & a \\ 2342+82 & 0.74 & 0.66 & 10-100 & 0.79 & 4.5 & $\la$1.3 & 0.82 & a \\ \hline B3\,VLA & & & & & & & & \\ 0809+404 & 0.55 & 4.80 & 4.8 & 0.53 & 1.1 & 16.1 & 0.48 & ? \\ 0810+460B & 0.33 & 2.80 & 13.9 & 0.94 & 1.1 & 10.3 & 0.44 & a \\ 1025+390B & 0.361& 5.10 &$>$100\phantom{.0}& 0.65 & 0.8 & 6.0 & 1.40 & a? \\ 1128+455 & 0.40 & 1.90 & 1.8 & 0.53 & 1.5 & 17.3 & 0.18 & ? \\ 1159+395 & 2.37 & 0.35 & 2.8 & 0.38 & 4.7 & 1.6 & 0.35 & ? \\ 1225+442 & 0.22 & 0.50 & 2.5 & 0.67 & 0.8 & 40.3 & 0.02 & ? \\ 1233+418 & 0.25 & 2.70 & 3.5 & 0.51 & 0.6 & 52.7 & 0.08 & a? \\ 1242+410 & 0.811& 0.40 & 4.7 & 0.38 & 2.6 & 4.2 & 0.16 & ? \\ 1350+432 & 2.149& 8.00 & 0.2 & 0.84 & 1.2 & 55.4 & 0.23 & ? \\ \hline \end{tabular} \end{table*} \begin{figure}[t] \epsfxsize=8cm\epsfbox [45 150 540 720] {8461.f4} \caption[]{$\chi^{2}_{\rm red}$ histograms for the CI, CIE, and CIm models. On the bottom right the histogram of the ratio between the CIm and the CI $\chi^{2}_{\rm red}$ is also shown.} \end{figure} \subsection{Radiative ages and the nature of CSS sources} In order to determine radiative ages, from Eq. (\ref{b}) and variants of the other models, the magnetic field $B$ has to be known. We stress that the age depends rather strongly on $B$, which is somewhat uncertain. We take the equipartition value $B_{\rm eq}$. Our assumption is motivated by the fact that $B_{\rm eq}$ accounts rather well for the low-frequency turnover in terms of synchrotron self-absorption. We are aware that this is not a proof for equipartition. Other authors (e.g. Bicknell et al. 1997) prefer instead thermal absorption. In any case, $B_{\rm eq}$ represents a poor statistical upper limit to $B$, in the sense that, were it larger by a factor of four, the low frequency turnovers would be systematically higher than observed by $\approx$ 30\%. Using the value of $B_{\rm eq}$, we obtain, from Eq. (\ref{b}), radiative ages $\tau_{\rm syn}$ ranging from $10^3$ to $10^5$ years. Since the intrinsic magnetic fields of the CSS sources in our sample strongly overweight the magnetic field equivalent to the cosmic microwave background ($B_{\rm CMB}$), the latter can be neglected in Eq. (\ref{b}). Therefore, if the source magnetic field deviates by a factor {\it f} from the field determined for equipartition $B_{\rm eq}$, the radiative ages will change by $f^{-3/2}$. These radiative ages do not necessarily represent the source ages, but rather the radiative ages of the dominant source component(s). Only when the lobes, which have accumulated the electrons produced over the source lifetime, dominate the source spectrum, the radiative age $\tau_{\rm syn}$ is likely to represent the age of the source. If, instead, the spectrum is dominated by a jet or by hot spots, the radiative age likely represents the permanence time of the electrons in that component and is expected to be less (perhaps much less) than the source age. In addition, dominant jets or hot spots might have their break frequency up-shifted by Doppler effects. The existing structure information on our sample, mostly from MERLIN and VLBI observations, allows us to split the sources in two groups: those in which the overall spectrum is dominated by lobes (classified ``type a'' in Tab.~2); those in which the spectrum is dominated by a bright jet or hot spot (classified ``type b'' in Tab.~2). For the B3\,VLA sample the available information does not yet allow such a morphological sub-division. The sources of class b have radiative ages systematically lower than those of class a, as we expected. Furthermore, while the synchrotron age is well correlated with the source size for class a, it seems that there is no correlation at all between the linear size and the radiative age for class b sources (see Fig.~5). \begin{figure}[t] \epsfxsize=8cm\epsfbox [45 150 540 680] {8461.f5} \caption[]{Linear size as a function of the synchrotron age for type a (filled dots) and type b (open dots) sources. The B3\,VLA sources have been excluded. The horizontal dashed lines represent the selection limits of the linear size distribution of the sources in our samples while the diagonal lines reflect constant values of $\frac{v_{\rm exp}}{c}$.} \end{figure} We have further computed for each source the expansion velocity $v_{\rm exp} \approx {\rm LS}/2 \tau_{\rm syn}$, where LS is the source largest dimension. The two classes show very different distributions:\\ Class a $\langle v_{\rm exp}/c \rangle = 0.34 \pm 0.07~~~ \sigma_{v_{\rm exp}} = 0.28$\\ Class b $\langle v_{\rm exp}/c \rangle = 11 \pm 4.4~~~ \sigma_{v_{\rm exp}} = 20$\\ Provided the assumed magnetic field is reasonably correct, the above values for the radiative ages indicate that {\it the CSS sources are young}. We further note that the ages, and corresponding expansion velocities, are not far (somewhat larger) from the recent results on expansion of CSOs by Owsianik et al. (1998) and Owsianik \& Conway (1998). As the radiative ages are strongly dependent on the assumed magnetic field, a field only a factor two lower than assumed would be required for a better agreement. \par \noindent In order to maintain the {\it frustration scenario}, in which the sources' lifetimes are $\approx 10^7$ years, their equipartition magnetic field should be decreased by a factor $\ga 20$. \par \noindent One could think that the correlation between linear sizes and the radiative ages shown in Fig.~5 could be a partial consequence of the equipartition assumption. In fact, $B_{\rm eq} \propto {\rm LS}^{-6/7} $ implies that ${\rm LS} \propto \tau_{\rm syn}^{7/9}$. This is not the case for the following reasons: 1) no correlation between linear sizes and radiative ages is obvious for type b sources. \par \noindent 2) In particular, the break frequencies seem to be correlated with the linear sizes for type a sources (Fig.~6). \par \noindent These means, at least for class a sources, that the break frequency is an effective clock indicating the source age. The correlation seen in Fig.~5 is not an artifact completely introduced by the assumption of equipartition in Eq. (2). \par \noindent \begin{figure}[t] \epsfxsize=8cm \epsfbox [45 150 540 680]{8461.f6} \caption[]{Linear size as a function of the break frequency (in the source rest frame) for type a (filled dots) and type b (open dots) sources. The B3\,VLA sources have been excluded. The horizontal dashed lines represent the selection limits of the linear size distribution of the sources in our samples.} \end{figure} \section{Conclusions} The high-frequency integrated spectrum of CSS sources shows a break with a moderate spectral steepening ($\Delta \alpha \approx 0.5$), well fitted by the continuous injection model spectrum with constant magnetic field. Spectral fits with synchrotron models with decreasing magnetic field are definitely poorer. There is no evidence of sharp cut-offs as would be expected if the supply of relativistic electrons had stopped more than $\approx 10^4$ years ago. In lobe dominated CSS sources the radiative ages $\tau_{\rm syn}$ are likely to represent the source ages. Assuming equipartition magnetic field strengths the source ages are in the range of up to $10^5$ years, in agreement with the recent results on expansion velocities of small size CSOs by Owsianik et al. (1998) and Owsianik \& Conway (1998) derived from VLBI observations. Therefore, assuming the CSS sources to be young, the magnetic field has been deduced to be equal to the equipartition value within a factor $\approx 2$. In jet or hot spot dominated CSS sources, the radiative life times are much shorter and very likely represent the permanence time of the electrons in those components. \begin{acknowledgements} We are grateful to Drs. B. Cotton and S. Spangler for very helpful comments on the manuscript of this paper and to Dr. H. Andernach for the suggestions on the compilation of the flux densities. We made use of the database CATS (Verkhodanov et al. 1997) of the Special Astrophysical Observatory. We acknowledge the Italian Ministry for University and Scientific Research (MURST) for partial financial support (grant Cofin98-02-32). Part of this work was supported by European Commission, TMR Programme, Research Network Contract ERBFMRXCT97-0034 ``CERES''. \end{acknowledgements} \newpage
\section{introduction} $SU(N)$ Yang-Mills (YM) theories exhibit peculiar and interesting features in 1+1 dimensions ($D=2$). The reduction from four to lower dimensions entails indeed tremendous simplifications, so that many problems can be faced, and often exactly solved \cite{wit},\cite{doug},\cite{basgri}. For instance exact evaluations of vacuum to vacuum amplitudes of Wilson loop operators, that, for a suitable choice of contour and in a particular limit, provide the potential between a static ${\rm q} {\rm \bar q}$ pair \cite{poly72},\cite{fish},\cite{wils74}, can be obtained. YM theories without fermions in 1+1 dimensions are considered free theories, apart from topological effects. This feature looks apparent when choosing an axial gauge. However, either when matter fields are introduced, or in Wilson Loop calculations, the perturbative $1+1$ dimensional theory exhibits dramatic infrared (IR) singularities which need to be regularized. Unfortunately the results appear to be dependent on such regularization procedures, even when they concern gauge invariant quantities \cite{tellu}. In light cone gauge (LCG) the (IR) singular behaviour is particularly apparent in the vector propagator, where the gauge pole conspires with the usual Feynman singularity to produce a double pole \cite{thooft}. A Cauchy principal value (CPV) prescription for this IR singularity has often been advocated \cite{call76}. It emerges quite naturally if the theory is quantized on the light cone surface $x^+=0$ \cite{Bas4}. On the other hand such a recipe is at odds with Wick's rotation. In ref.\cite{Wu} a causal prescription for the double pole has been proposed, which is nothing but the one suggested years later by Mandelstam and Leibbrandt (ML) \cite{ML}, when restricted to $1+1$ dimensions. This prescription follows from equal-time quantization \cite{Bas5} and is mandatory in order to renormalize the theory in 1+3 dimensions\cite{Bas3},\cite{Bas4}. In view of the above-mentioned results and of the fact that ``pure'' YM theories do not immediately look free in Feynman gauge, a systematic investigation has been undertaken to clarify their properties when the two dimensional picture is reached starting from higher dimensions. Since no exact solutions are available beyond $D=2$, the investigation has been focussed on perturbative calculations, looking for consistency checks, in particular testing the gauge invariance of the theory which holds order by order in the coupling constant expansion. Recalling that perturbative S-matrix elements cannot be consistently defined in non-Abelian gauge theories, owing to their (IR) singular mass-shell behaviour, the natural gauge invariant quantities to be considered are Wilson loops. A first test of gauge invariance in $1+3$ dimensions has been performed in refs.\cite{Korc,Bas2} by calculating at ${\cal O}(g^4)$, both in Feynman and in light-cone gauge with ML prescription, a rectangular Wilson loop with light-like sides, directed along the vectors $n_\mu = (T, - T)$, $n^*_\mu = (L, L)$ and parametrized according to the equations: \begin{eqnarray} \label{uno} C_1 &:& x^\mu (t) = n^{* \mu} t,\nonumber\\ C_2 &:& x^\mu (t) = n^{* \mu} + n^\mu t,\nonumber\\ C_3 &:& x^\mu (t) = n^\mu + n^{* \mu}( 1-t), \nonumber\\ C_4 &:& x^\mu (t) = n^\mu (1 - t), \qquad 0 \leq t \leq 1. \end{eqnarray} In order to perform the test, dimensional regularization ($D=2\omega$) was used for both UV and IR singularities. Full consistency between Feynman and light-cone gauge with the ML prescription was obtained. Since results in $2\omega$ dimensions were available, in view of the peculiar features of Yang-Mills theories in 2 dimensions mentioned above, the interest arose in knowing the outcome of the check in the limit $\omega \to 1$. The following unexpected results were obtained in \cite{Bas1}. The ${\cal O}(g^4)$ perturbative loop expression in $d= 1+(D - 1)$ dimensions is {\it finite} in the limit $D\to 2$. The loop expression is a function only of the area $n\cdot n^{*}$ for any dimension $D$ and exhibits also a dependence on $C_A$, the Casimir constant of the adjoint representation. In LCG this dependence comes from two sources: \begin{itemize} \item diagrams with two crossed propagators (colour factor $C_F(C_F - C_A/2)$, $C_F$ being the Casimir constant of the fundamental representation); \item a genuine contribution to the Wilson loop proportional to $C_F C_A$ coming from the one-loop correction to the vector propagator (self-energy diagram). \end{itemize} We shall concentrate our interest on the contribution due to this self-energy diagram. At a first sight, it is surprising, since, in 1+1 dimensions, there is no triple vector vertex in axial gauges. What happens is that the vanishing strength of the vertex at $D=2$ matches the self-energy loop singularity, eventually producing a finite result. Feynman diagrams with a triple vertex but no loops tend instead smoothly to zero when inserted in the Wilson contour. We notice that no ambiguity affects the ${\cal O}(g^4)$ gauge invariant result, which is finite; in addition the presence of $C_A$ cannot be re-absorbed by a redefinition of the coupling, that, while unjustified on general grounds, would also turn out to be dependent on the area of the loop. In order to clarify whether the appearance of $C_A$ in the maximally non-Abelian term is indeed a pathology, one should examine the potential $V(2L)$ between a ``static" ${\rm q} {\rm\bar q}$ pair in the fundamental representation, separated by a distance $2L$. Therefore in ref.\cite{bello} we have considered a different Wilson loop, {\it viz} a rectangular loop with one side along the space direction and one side along the time direction, of length $2L$ and $2T$ respectively. Eventually the limit $T \to \infty$ at fixed $L$ is to be taken: the potential $V(2L)$ between the quark and the antiquark is indeed related to the value of the corresponding Wilson loop amplitude ${\cal W}(L,T)$ through the equation \cite{ALTE} \begin{equation} \lim_{T\to\infty}{\cal W}(L,T)=const.\ e^{-2i T V(2L)}\ . \label{potential} \end{equation} The crucial point to notice in eq.(\ref{potential}) is that dependence on the Casimir constant $C_A$ should cancel at the leading order when $T\to \infty$ in any coefficient of a perturbative expansion of the potential with respect to coupling constant. This criterion has often been used as a check of gauge invariance \cite{Bas4}. In ref.\cite{bello} the calculation has been performed in Feynman gauge, obtaining the following results. For $D>2$ the ${\cal O}(g^4)$ perturbative expression of the loop depends, besides on the area, also on the ratio $\beta=L/T$. As we are eventually interested in the large-$T$ behavior, we have always considered the region $\beta<1$; moreover we have chosen $D=2+\epsilon $ with a small $\epsilon >0.$ As long as $D>2$, agreement with Abelian-like time exponentiation (ALTE) occurs in the limit $T\to \infty$, with a pure $C_F$-dependence in the leading coefficient. Consistency of all previous results \cite{Bas4} in higher dimensions is thus re-established. The limit $D\to 2$ for $\beta=0$ {\it exactly} reproduces the gauge invariant result obtained in ref.\cite{Bas1} for a loop of the same area with light-like sides; thereby we enforce the argument that in two dimensions a pure area behaviour is expected, no matter the orientation and the shape of the loop. What may be surprising is that the term, which in LCG corresponds to the self-energy correction, exhibits, in the limit, a pure area dependence on its own. However, in two dimensions at ${\cal O}(g^4)$, a $C_A$-dependence is definetely there and agreement with ALTE is lost. Actually this behaviour at $D=2$ persists at any order of $g$ and affects the sum of the perturbative series \cite{stau},\cite{basgri}. A peculiar feature of the light-cone gauge in $2$ dimensions is that individual Wilson loop diagrams do not exhibit any singularity; hence there is no need of dimensional regularization. In ref. \cite{Bas7}, a ${\cal O}(g^4)$ perturbative calculation of the Wilson loop in LCG with ML prescription, for a rectangular loop with sides $2T\times 2L$ lying in the $x^0\times x^1$ axes, was performed at $D=2$. No agreement occurs with the result one finds in ref.\cite{bello} when taking the limit $D\to 2$. The source of such a discrepancy is rooted in the mentioned self-energy diagram contribution, which is obviously missing at $D=2$, but provides a finite term in the limit $D\to 2$, thereby producing a discontinuity in the theory \cite{Bas1}. The purpose of this paper is to check explicitly this property by evaluating in LCG the relevant discontinuity for the Wilson loop of ref. \cite{Bas7}. We confirm that the missing term comes from the diagram with a self-energy corrected propagator, evaluated at $D=2+\epsilon$, when eventually taking the limit $\epsilon \to 0$. We thereby reproduce for a space-time contour the phenomenon in LCG found in ref.\cite{Bas1} for a contour with light-like sides. Actually, from the computation of the self-energy diagram at $D>2$, we find, as an extra bonus, that its contribution vanishes for $\epsilon >0$ in the limit $T\to \infty$ with the same ``universal'' factor $T^{4-4\omega}$ we have obtained in ref.\cite{bello} for the maximally non-Abelian contributions \cite{tay}. The limits $T \to \infty$ and $\epsilon \to 0$ {\it do not commute}. \section{The calculation} We recall some basic notions and notations. We consider, as in ref.\cite{bello}, the closed path $\gamma$ parametrized by the following four segments $\gamma_i$, \begin{eqnarray} \gamma_1 &:& \gamma_1^\mu (s) = (sT, L)\ ,\nonumber\\ \gamma_2 &:& \gamma_2^\mu (s) = (T,-sL)\ ,\nonumber\\ \gamma_3 &:& \gamma_3^\mu (s) = (-sT, -L)\ , \nonumber\\ \gamma_4 &:& \gamma_4^\mu (s) = (-T, sL)\ , \ \ \qquad -1 \leq s \leq 1. \label{path} \end{eqnarray} describing a (counterclockwise-oriented) rectangle centered at the origin of the plane ($x^1,x^0$), with length sides $(2L,2T)$, respectively. The perturbative expansion of the Wilson loop is \begin{equation} {\cal W}_\gamma (L,T) = 1 + {1\over N}\sum_{n=2}^\infty (ig)^n \oint_\gamma dx_1^{\mu_1} \cdots \oint_\gamma dx_n^{\mu_n}\theta( x_1 >\cdots >x_n ) {\rm Tr} [ G_{\mu_1 \cdots \mu_n} (x_1,\cdots ,x_n)]\ , \label{wilpert} \end{equation} where $ G_{\mu_1 \cdots \mu_n} (x_1,\cdots ,x_n)$ is the Lie algebra valued $n$-point Green function, and the Heavyside $\theta$-functions order the points $x_1,\cdots ,x_n$ along the integration path $\gamma$. It is easy to show that the perturbative expansion of ${\cal W}_\gamma$ is an even power series in the coupling constant, so that we can write \begin{equation} {\cal W}_\gamma (L,T)= 1+g^2 {\cal W}_2 + g^4 {\cal W}_4 + {\cal O}(g^6)\ . \label{pert} \end{equation} To have a sensitive check of gauge invariance, one has to consider at least the order $g^4$, ({\it i.e.} one has to evaluate ${\cal W}_4$), as this is the lowest order where genuinely non-Abelian $C_FC_A$ contributions may appear. In turn, in the calculation of ${\cal W}_4$, only the so called maximally non-Abelian contribution ${\cal W}_4^{na}$ need to be evaluated, that in our case comes from the terms proportional to $C_FC_A$. The Abelian contribution, proportional to $C_F^2$, can be easily obtained thanks to the Abelian exponentiation theorem \cite{tay}. The diagrams contributing to ${\cal W}_4^{na}$ can be grouped into three families: a) crossed diagrams (${\cal C}_{(ij)(kl)}$), with a double gluon exchange in which the two (crossed) propagators join the sides $(ij)$ and $(kl)$ of the contour $\gamma$; b) spider diagrams (${\cal S}_{ijk}$), which are obtained by attaching a three point Green function at the tree level to the sides $(ijk)$ of the loop; c) bubble diagrams (${\cal B}_{ij}$) , that are single exchange diagrams in which the gluon propagator, corrected by a self-energy term, joins the sides $(ij)$ of the contour. In arbitrary dimensions, the calculation of the Wilson loop is much more awkward in LCG than in covariant gauge, due to a more complicated form of the vector propagator. However, when considering the $D\to 2$ limit, diagrams in LCG have much better analyticity properties in $\omega$ than the ones in Feynman gauge. The vector propagator in LCG with ML prescription is a tempered distribution at $D=2$, at odds with the one in Feynman gauge. Moreover it is summable along the (compact) loop contour. Due to this property, we can conclude that all the maximally non-Abelian contributions arising from diagrams with crossed propagators sum to an expression that, in the limit $D\to 2$, reproduces the result of ref. \cite{Bas7}, namely \begin{equation} \label{croci} {\cal W}^{cr}=C_AC_F{(LT)^2\over 3}\ \ . \end{equation} \smallskip Now we consider the contribution ${\cal W}^{bub}$ coming from bubble diagrams. In LCG and on the plane $x^0\times x^1$, the only non-vanishing component of the two point Green function $\Delta_{\mu\nu}$ at the order ${\cal O}(g^2)$ is $ \Delta_{++}(x)\equiv \Delta(x)$, that reads, at $x_\perp=0$ \cite{Bas2}, \begin{equation} \label{self} \Delta(x)=-\frac{g^2}{8\pi^{2\omega}}C_A\frac{(x^-)^2}{(-x^2+i\varepsilon)^{2\omega-2}} f(\omega)\ , \end{equation} \begin{equation} \label{fd} f(\omega)=\frac{1}{(2-\omega)^3}\left[\frac{\Gamma^2(3-\omega)\Gamma(2\omega -3)}{\Gamma(5-2\omega)} -{\Gamma(\omega-1) \Gamma(\omega) (10\omega^2 -19\omega + 10)\over 4(2\omega-3)(2\omega-1)}\right] \ . \end{equation} Following the notations of ref. \cite{bello}, there are 10 topologically inequivalent bubble diagrams. However, due to the symmetry of the Green function and to the symmetric choice of the contour, only six of them are independent, and the ${\cal O}(g^4)$ contribution to the Wilson loop arising from bubble diagrams can be written as \begin{equation} \label{bub1} {\cal W}^{bub}= 2({\cal B}_{11} + {\cal B}_{22} +{\cal B}_{13}+ {\cal B}_{24} + 2{\cal B}_{12}+ 2{\cal B}_{14}) \ , \end{equation} where each single contribution ${\cal B}_{ij}$ can be calculated by replacing eqs. (\ref{path}), (\ref{self}) in the formula \begin{equation} \label{bubbola} {\cal B}_{ij}=-{1\over 2}g^2 C_F \int_{-1}^1 ds \int_{-1}^1 dt \Delta_{\mu\nu} (\gamma_i (s)- \gamma_j(t)) \dot \gamma^\mu_i (s)\dot \gamma^\nu_j (t)\ , \end{equation} where the dot denotes derivative with respect to the variable parametrizing the segment. The calculation being standard, we shall report only the final result \begin{eqnarray} \label{bub2} &&{\cal W}^{bub}= {C_FC_A\over \pi^{2\omega} }f(\omega) (LT)^2 (2L)^{4-4\omega}\left\{e^{-2i\pi\omega }\beta^{4\omega-6}\left[ {1\over (7-4\omega)(8-4\omega)}\right.\right.\nonumber\\ &&\times\biggl(1-(8-4\omega) _2F_1(2\omega-2,2\omega-7/2;2\omega-5/2;\beta^2)+(7-4\omega) (1-\beta^2)^{3-2\omega}\biggr)\nonumber\\ && \left. -{1\over(3-2\omega)(4-2\omega)}\left(1-(1-\beta^2)^{4-2\omega}\right) + {5-2\omega\over(6-4\omega)(4-2\omega)} \left(1-(1-\beta^2)^{3-2\omega}\right) \right]\nonumber\\ &&+e^{-2i\pi\omega} \beta^{4\omega-4}\left[{(1-\beta^2)^{3-2\omega} \over (3-2\omega)(4-2\omega)}- { _2F_1(2\omega-2, 2\omega-5/2;2\omega-3/2;\beta^2)\over (5-4\omega)}\right.\nonumber\\ &&\left. - _2F_1(2\omega-2,1/2;3/2;\beta^2){ { } \over { } }\right] +i\beta {\sqrt{\pi} (\omega-2)\Gamma(2\omega -7/2)\over \Gamma(2\omega -2)} \nonumber\\ &&\left. -e^{-2i\pi\omega} {\beta^{4\omega-2}\over 3} {_2F_1}(2\omega-2,3/2;5/2;\beta^2) + { \beta^2\over (7-4\omega)}\right\}\ , \end{eqnarray} where $\beta = L/T$. Some comments are here in order. First of all there is a dependence on the dimensionless ratio $\beta$, besides the area, at variance with the analogous result in LCG for the rectangle of light-like sides. However, in the equation above, one can easily check that the quantity ${\cal W}^{bub}/(LT)^2$ is not singular for $\beta\to 0$. Actually eq.(\ref{bub2}) exhibits, for $\omega >1$, the expected damping factor $T^{4-4\omega}$ in the large-$T$ limit. In the limit $\omega\to 1$ the dependence on $\beta$ disappears and the pure area law is recovered: ${\cal W}^{bub} =C_FC_A (LT/\pi)^2$. This is exactly the ``missing'' term to be added to the expression of ref.\cite {Bas7} to obtain the final result for the maximally non-Abelian contribution to the perturbative ${\cal O}(g^4)$ Wilson loop in the limit $D\to 2$, \begin{equation} \label{finale} {\cal W}_4^{na} = C_FC_A \left({LT\over \pi}\right)^2 \left[ 1 + {\pi^2\over 3}\right]\ . \end{equation} Equation (\ref{finale}) is in full agreement not only with ref. \cite{bello}, where an anologous Wilson loop was calculated in Feynman gauge, but also with ref. \cite{Bas1}, where the loop was oriented in a different direction. Moreover, in LCG, different families of diagrams (``crossed'' and ``bubble'' diagrams) give the same contribution ($C_FC_A {(LT)^2 \over 3}$ and $C_FC_A \left({LT\over \pi}\right)^2$ respectively) no matter the orientation of the loop: remarkably, invariance under area-preserving diffeomorphisms is recovered in the limit $D\to 2$, even when the Wilson loop is first evaluated in higher dimensions, and then the limit $D\to 2$ is taken. In turn the result above implies that ``spider'' diagrams, namely diagrams with a triple vector vertex, cannot contribute in the limit $D \to 2$. This is not surprising, as the same phenomenon occurred in ref. \cite{Bas1}, although for a different contour (contour with light-like sides). In order to support this conclusion, we show that the relevant three point Green function at ${\cal O}(g)$, vanishes when $D \to 2$. To this aim, let us consider the three point Green function ${\cal V}_{\mu\nu\rho} (x,y,z)$. Due to the LCG choice, its only non-vanishing component when considering the loop in the $x^0\times x^1$ plane is ${\cal V}(x,y,z) = {\cal V}_{+++}(x,y,z)$; up to an irrelevant multiplicative constant, it is given by \begin{equation} \label{gamma} {\cal V}(x,y,z)=\int d^{2\omega}\zeta {\partial\over \partial z^\alpha} \left[{\partial\over \partial x^\alpha}{\partial\over \partial y^+}- {\partial\over \partial y^\alpha}{\partial\over \partial x^+}\right] F(x-\zeta)F(y-\zeta) G(z-\zeta) \end{equation} $$ +{\rm cycl.\ perm.} \ \{x,y,z\}\ \equiv ({\cal V}_1-{\cal V}_2) +{\rm cycl.\ perm.} \ \{x,y,z\}\ .$$ Here the index $\alpha$ runs over the transverse components and the functions $G$ and $F$ are the following Fourier transforms \begin{equation} \label{g} G(x)=\int d^{2\omega} p {e^{ipx}\over p^2 +i\varepsilon}=-\pi^\omega \Gamma(\omega-1) \left(-{x^2\over 4} + i\varepsilon\right)^{1-\omega}, \end{equation} \begin{equation} \label{f} F(x)=\int d^{2\omega} p {e^{ipx}\over (p^2 +i\varepsilon) (p^++i\varepsilon p^-)}= -i\pi^\omega \Gamma(\omega-1)\int_0^{x_+} d\rho \left({x_\perp^2 - 2 x_- \rho\over 4} + i\varepsilon\right)^{1-\omega}. \end{equation} Let us consider, for instance, the first term in eq. (\ref{gamma}), that we call ${\cal V}_1$. Using standard Feynman integrals techniques, integrations over momenta and over the intermediate point $\zeta$ can be performed, so that ${\cal V}_1$ can be rewritten, after some convenient change of variables, as \begin{eqnarray} \label{gamma1} {\cal V}_1&=&{i\pi^\omega (4 \pi)^{3\omega}\over 8}\Gamma(2\omega-1) (\omega-1)\!\! \int_0^1 \!\! d\xi d\eta d\mu \, \eta [\mu(1-\mu)]^{\omega -1} \int_0^\infty \!\! d\tau {[1+\tau (\mu\xi + \eta (1-\mu))]^{2\omega-5}\over (1+\tau)^{\omega}}\nonumber \\ &&\times {[(x-z)_+ +\tau\eta (1-\mu) (x-y)_+][(y-z)_+ + \tau \mu\xi (y-x)_+]^2\over[-\mu\xi(x-z)^2 -\eta (1-\mu) (y-z)^2 - \tau\xi\eta\mu(1-\mu) (x-y)^2 +i\varepsilon]^{2\omega-1}}\ . \end{eqnarray} Since ${\cal V}_1$ has an explicit zero at $\omega=1$, if we show that the integral in (\ref{gamma1}) is convergent when evaluated at $\omega=1$, we have proved that the three point Green function vanishes at $D=2$. Integral (\ref{gamma1}) is discussed in the Appendix. \section{Conclusions} A peculiar feature of the light-cone gauge formulation of Yang-Mills theories is that they can be consistently defined in two dimensions: contrary to the covariant Feynman gauge, the light-cone gauge propagator with ML prescription for the spurious pole is a tempered distribution at $D=2$. In particular, the large $T$ behaviour of the Wilson loop can be evaluated without the need of introducing any regulator; the finite result has been presented in ref.\cite{Bas7}. This result, however, cannot be compared with the result one would obtain in Feynman gauge, as in the latter case, the free propagator is not a tempered distribution at $D=2$. In Feynman gauge the best one can do is to evaluate the Wilson loop in $D$ dimensions, and to take eventually the limit $D\to 2$. In so doing one obtains again a finite result \cite{bello} that, however, is {\it different} from the one of ref.\cite{Bas7}. In LCG the diagram with a self-energy correction in the propagator, which only exists in $D>2$, makes the difference. It is precisely the contribution we have evaluated in this paper. It provides us with the missing term to get agreement between refs. \cite {Bas7} and \cite{bello}, {\it i.e.} to recover gauge invariance. Such a phenomenon was not unexpected in the light of ref.\cite{Bas1}. Perturbative Yang-Mills theory in LCG looks indeed discontinuous in the limit $D\to 2$; actually, starting from a vanishing coupling at $D=2$, it exhibits a kind of ``instability'' with respect to a change of dimensions. On one hand our result clarifies the nature of the discontinuity of Yang--Mills theories in two dimensions, on the other it rises new interesting questions for future investigations. While in any dimension $D>2$ $perturbative$ Wilson loop calculations are in agreement with Abelian-like time exponentiation, as all $C_A$ dependent terms turn out to be depressed in the large-$T$ limit, at $D=2$ neither the result in ref. \cite{Bas7} nor the one in \cite{bello} share this property, as they both exhibit an explicit $C_A$-dependence in the coefficient of the leading term when $T \to \infty$. At $D=2$ exponentiation in terms of $C_F$ occurs perturbatively only in light-front formulation (ref. \cite{thooft}); in equal-time quantization, exponentiation requires full resummation of genuine non-perturbative contributions (instantons) \cite{basgri}. The difference between the formulations above (and their related vacua) as well as the reason why this phenomenon seems to be crucial only at $D=2$ are under active investigation.
\section{INTRODUCTION} In the past few years it has been shown that the Skyrme model \cite{skyrme} is related to the low energy limit of QCD \cite{witten:83}. This fact, together with progress in the approximate and numerical solution of strongly nonlinear equations, has renewed interest in a detailed study of the quantum mechanics of the Skyrme model. These studies are directed along two main lines. The first simplifies the model at the classical level to the so-called Baby Skyrmion model, which in turn is related to the Sine-Gordon equation \cite{skyrme,KPZ,PZ}. In this simplified classical model extensive numerical studies have led to an increased understanding of the stability, scattering and interaction of coherent structures with radiation \cite{KPZ,PZ}. On the other hand, the second line has focused on the quantum effects of the full Skyrme model \cite{adkins,gisiger}. In these studies the quantization was either obtained by linearization around a field configuration \cite{adkins} or by a finite dimensional approximation to the full problem \cite{gisiger}. Here we take a complementary approach. We study the 1+1 Sine-Gordon equation keeping all the degrees of freedom and quantize along the lines of \cite{eboli1,eboli2,ours,cooper}. This leads to a field equation which is strongly coupled with the equations for the fluctuations. In \cite{cooper} the formalism for the functional coherent state approximation was fully developed and the possible advantages and disadvantages of various approaches designed for numerical purposes were discussed in detail. In particular, the closed-time path method with the Hartree factorization (see e.g. \cite{keldish}) was applied in \cite{cooper} to the static Sine-Gordon field, and the approximate results obtained for phase transitions and stability were found to compare favorably with known exact results. In this same work static results were also obtained for more realistic field equations. Our present work differs from the above cited papers in that we choose to approximate the Green function (the variance kernel for the Gaussian ansatz) by a suitable parametrized trial function. This choice leads to a great simplification of the problem for the case of low momentum fluctuations. Thus, for example, the infinite system of partial differential equations of \cite{eboli1} become ordinary differential equations and, eventhough the field and the fluctuations in our formalism are strongly coupled, we are able to use some of the solutions of the classical Sine-Gordon equation (since in $1+1$ dimensions it is completely integrable) to construct approximate solutions to the quantum problem, including the effect of the radiation generated by the quantum fluctuations. We then solve the equations numerically and these numerical solutions are compared with asymptotic solutions. Note also that the approach used in the present work is completely different to that of \cite{lu,lu2} where the wave functionals are constructed using Gaussian approximations to the functional Schr\"odinger equation for the Sine-Gordon field. However, the particles are considered as higher excited states (in function space) of the linearized field equations. In our treatment the field equations are nonlinear and dynamic and different particles are represented by different nonlinear field configurations, not by higher order Hermite functionals as in \cite{lu,lu2}. Finally, it must be stressed that the approximate analytic results obtained here are valid in a strongly nonlinear regime and, in principle, do not depend on the $1+1$ nature of the model and could be used to study problems related to more realistic simplifications of the Skyrme model. The paper is organized as follows. In Section II the detailed formulation of the quantum problem is stated, with the free parameters adjusted to mimic mesons and baryons. Section III is devoted to the study of the coherent state approximation and the derivation of the quantum equations for the field for low momentum fluctuations. In Section IV three problems are considered. The first is the nonlinear stability of a single Soliton under the influence of quantum fluctuations. This stability is studied both numerically and asymptotically. In particular the asymptotic solution includes the damping effect of the radiation shed by the Soliton due to the fluctuations. This asymptotic solution explains in detail the mechanism for the nonlinear stability of the Soliton. The second problem studied is the scattering of a meson (wave) by a static Soliton. The numerical solution for this problem shows that in this process the Soliton is also stable. Finally the third problem studied concerns the collision of Solitons and the quantum evolution of a bound state (Soliton and an anti-Soliton). It is shown that the complete integrability of the Sine-Gordon simplification of the full Skyrme model to just one internal degree of freedom for the field precludes the processes of fusion and splitting. The processes of fusion and splitting are produced by the influence of all the internal degrees of freedom. \section {FORMULATION OF THE PROBLEM} For the basic structure we take the baby Skyrme model, which is a reduction of the full model with only two fields present \cite{skyrme,KPZ}. This model, in turn, reduces to the Sine-Gordon equation which is known to be completely integrable \cite{KPZ}. In these variables the Hamiltonian takes the form \begin{equation} H = m {c}^2 l \int {\left[ \frac{1}{2} \pi^2 + \frac{1}{2} \left( \frac{\partial\varphi}{\partial x} \right)^2 + l^{-2} (1 - \cos \varphi)\right] dx}, \label{e:ham} \end{equation} where $\varphi$ is an angle variable whose momentum $\pi$ is given by \begin{equation} \pi =\frac{1}{c} \frac{\partial\varphi}{\partial t}, \label{e:pi} \end{equation} $m$ is the mass of the particle and $l$ is a typical particle size. Using dimensionless variables $\tilde x = x/l $ and $ \tilde t = ct/l$, we obtain, after dropping the tildes, \begin{equation} H = m {c}^2 \int {\left[ \frac{1}{2} \pi^2 + \frac{1}{2} \left( \frac{\partial\varphi}{\partial x} \right) ^2 + (1 - \cos \varphi)\right] \: dx}, \label{e:nondimham} \end{equation} with $\pi = \partial\varphi / \partial t$. To mimic a baryon by means of the Sine-Gordon soliton, we take $l \sim 10^{-13}$ cm and $ m \sim 10^{-27}$ Kg, which gives an internal time of $10^{-23}\rm {sec}$. The equation of motion derived from the Hamiltonian variational principle \begin{equation} \delta \int_{t_0}^{t_1} \left( \pi \dot{\varphi} - H \right) \: dt \label{e:varham} \end{equation} is the Sine-Gordon equation \begin{equation} \frac{\partial^2 \varphi}{\partial t^2} - \frac{\partial^2 \varphi}{\partial x^2} + \sin \varphi =0. \label{e:sineg} \end{equation} The one dimensional Skyrmion is the soliton solution \begin{equation} \varphi = - 4 \mathop{\rm arctan}\nolimits \left[ \exp \left( - (x - vt) / \sqrt{1-v^2} \right) \right], \label{e:sgsol} \end{equation} of \ref{e:sineg}, which represents a localized deformation at $x = vt$. In this context $v$ must be taken to satisfy $v \ll 1$, since the Skyrme model is only consistent for small energies. The linear travelling periodic wave solutions of the Sine-Gordon equation are interpreted as pions. Notice that the nonlinear model contains linear waves which describe bosons and nonlinear structures which describe fermions. In his original work Skyrme suggested that the fermionic part of the Lagrangian is needed only just to count the number of localized states of finite amplitude of the bosonic field \cite{skyrme}. In this interpretation, fermions are just a point approximation to nonlinear localized bose fields. It has also been suggested on rather general grounds that the canonical quantization of fields (as bosons) gives Fermi-Dirac type statistics for the kinks \cite{finkelstein}. In the Sine-Gordon model the exclusion principle holds for kinks, since we know that for the general exact solution there is no solution with two identical solitons \cite{lamb}. In this article, we shall consider a canonical quantization using the functional Schr\"odinger picture. It is to be noted that all quantizations for the Skyrme model cited in the Introduction make the same assumption. However, in this work we differ from previous treatments in that we shall keep infinitely many degrees of freedom in the classical field and reduce the dimensionality of the space of fluctuations. The final result will be shown to be a system which consists of a partial differential equation (similar to the Sine-Gordon equation) for the field coupled to a system of nonlinear ordinary differential equations which control the fluctuations. In the field configuration representation the Schr\"odinger equation takes the form \begin{equation} i \hbar \frac{\partial \Psi(\varphi)}{\partial t} = \hat H \Psi(\varphi), \label{e:5} \end{equation} where $\Psi(\varphi(x), t):= \langle \varphi(x) |\Psi(t)\rangle$ is the amplitude for finding the field system characterized by the state vector $|\Psi(t)\rangle$ in the field configuration $\varphi(x)$ at time $t$. In this configuration representation the scalar product of two state vectors is given by the functional integration \begin{equation} \langle \Psi_{1} | \Psi_{2} \rangle = \int {\Psi^\ast _{1}(\varphi,t)} \Psi_2 (\varphi,t) \: {\cal D} \varphi, \label{e:funint} \end{equation} and the field operators are represented by functional kernels. Thus the field operator $\hat \phi(x)$ is represented by $\langle \varphi(x) | \hat \phi(x)|\Psi(t)\rangle = \varphi(x) \Psi(\varphi(x),t)$ and, therefore, acts as a multiplication operator, while the action of the canonical momentum operator is given by \begin{equation} \langle \varphi(x)|\hat \pi (x')| \Psi(t)\rangle = -i \hbar \frac{\delta}{\delta \varphi(x')} \langle \varphi(x) |\Psi(t)\rangle . \label{e:conmom} \end{equation} The Hamiltonian field operator $\hat H$ then becomes \begin{equation} \hat H = mc{^2} \int \left[ \frac{1}{2} {\hat \pi}^2 (x) + \frac{1}{2} \left( \frac{\partial\hat \varphi}{\partial x} \right)^{2} + (1 - \cos \hat \varphi) \right] \: dx, \label{e:fieldop} \end{equation} where there is no ambiguity in the ordering and the functions of operators are defined by their power series. The quantum mechanical problem for the field consists of solving the Schr\"odinger equation (\ref{e:5}) for a given initial field configuration. Notice that the time in (\ref{e:5}) has a scale set by $mc^{2} / \hbar$, which is of the same order of magnitude ($ 10^{-23}$) as the time scale $c / l$ for the field fluctuations. This is to be expected since the Skyrme equation was assumed to be consistent with the quantum scale of the particle. Thus we can take the same dimensionless time variable for either scale. \section{APPROXIMATE SOLUTIONS TO THE FUNCTIONAL EQUATION} To solve the Schr\"odinger equation (\ref{e:5}), we take a coherent state approximation \cite{eboli1,eboli2,ours,cooper} and study the evolution of its parameters. Following \cite{eboli1,eboli2,ours,cooper} we consider the functional \begin{equation} \Gamma = \int \langle \Psi | i \frac{\partial}{\partial t} - \hat H | \Psi \rangle \: dt, \label{e:gamma} \end{equation} which we extremize with the Gaussian trial functional \begin{eqnarray} \Psi(\varphi(x) ,t)&=& \exp \left\{ i \int \pi (x,t) \left[ \varphi (x) - \phi (x,t)\right] \: dx \right\} \nonumber \\ & & \: \exp \left\{ - \int \int dx dy \: \left[ \varphi (x) - \phi (x,t) \right] \right. \nonumber \\ & & \left. (\times)\left[ \frac{1}{4} \Omega ^{-1} (x,y ,t) - i \Sigma (x,y,t) \right] \left[ \varphi (y) - \varphi (x) \right] \right\} , \label{e:trial} \end{eqnarray} where the kernels $\Omega^{-1}$ and $\Sigma$ take into account the quantum fluctuations and $\phi(x,t)$ and $\pi (x,t)$ are the average field and average momentum of the Gaussian, respectively. Substituting the trial function (\ref{e:trial}) into the functional (\ref{e:gamma}) and integrating over the $\varphi(x)$ variable yields an averaged action \cite{eboli1,eboli2,ours,cooper}. This action is given in terms of $\phi(x,t)$, $\Omega$ and $\Sigma$. The potential is then also expanded around the average $\phi(x,t)$. It should be stressed that the consistency of this approximation depends on the smallness at all times of the variance $\Omega$, in the sense that the average energy of the fluctuations around the mean $\phi(x,t)$ is small compared with the energy of the average motion. Moreover, by keeping only quadratic terms resulting from the averaging around the mean of the nonlinear term, which amounts to making the assumption that the energy of the fluctuations is small compared to the energy of the mean, we arrive at an effective action of the form \begin{eqnarray} L &=& \int_{0}^{T} \left\{ \int_{-L}^{L} \left[ \pi \frac{\partial \varphi}{\partial t} - \left( \frac{1}{2} \left(\frac{\partial \varphi} {\partial x}\right)^{2} + (1 - \cos \varphi )\right)\right. \right. \nonumber\\ & & \mbox{} + \Sigma {\bf\dot{\Omega}}(x,x,t) -2 \Sigma \Omega \Sigma (x,x,t) - \cos (\varphi) \Omega(x,x,t)\nonumber \\ & & \mbox{} \left. \left. + \frac{1}{2} \frac{\partial^{2} \Omega}{\partial x^{2}}(x,y,t)|_{x=y} - \frac{1}{8}\Omega^{-1}(x,x,t) \right] \right\} \: dxdt + O(\Omega^{2}), \label{e:effaction} \end{eqnarray} where the notation $\Omega \Sigma$ is taken to mean the kernel of the operator defined by the convolution of $\Sigma$ with $\Omega$, that is \begin{equation} \Omega \Sigma (x,y) = \int_{-L}^{L} \Omega(x,z) \Sigma(z,y) \: dz \label{e:conv} \end{equation} and $\Sigma \Omega \Sigma$ is the kernel of the operator defined by the convolution of $\Sigma$, $\Omega$ and $\Sigma$ \begin{equation} \Sigma \Omega \Sigma (x,y) = \int_{-L}^{L} \int_{-L}^{L} \Sigma (x,z) \Omega (z,u) \Sigma (u,y) \: du dz. \label{e:sos} \end{equation} In the following we will take the limit $L \rightarrow \infty$ at different stages in the calculation of the effective action. The effective action can be computed once we choose an appropriate parameterization for the variance $\Omega$ and the phase $\Sigma$. It must be noted that thanks to the simple form of the potential. the Gaussian integral can be evaluated exactly. However, since we are interested in small fluctuations we stop at the quadratic level. We will come back to this point when comparing our results with those in \cite{cooper}. Now, since the field is homogeneous, we can take \begin{eqnarray} \Omega (x,y,t)&=& \frac{1}{2 \pi} \int_{-\infty}^{\infty} \exp [ik(x-y)] \hat \Omega (k,t) \: dk \label{e:hom1} \\ \Sigma (x,y,t)&=& \frac{1}{2\pi}\int_{-\infty}^{\infty} \exp [ik(x-y)] \hat \Sigma (k,t) \: dk, \label{e:hom2} \end{eqnarray} where \begin{eqnarray} \hat \Omega (k,t)&=& \frac{\Omega_{0}}{k^{2} + a^{2}(t)}, \label{e:omhat} \\ \hat \Sigma (k,t)&=& \frac{b(t)}{k^{2} + a^{2}(t)}. \label{e:sighat} \end{eqnarray} The parameters $a(t)$ and $b(t)$, which control the spreading of the fluctuations, are to be determined after the effective action is varied. The choice of the trial function is then guided by the simplicity of the resulting expressions. Notice, however, that since the results obtained depend only on the spreading, the same qualitative behaviour will be obtained for other forms of the trial function. Since the approximate solution (\ref{e:trial}) involves the kernel $\Omega ^{-1}$ the proposed expression is convergent provided that the momentum of the fluctuations involved in the integration is low. This assumption is in agreement with the fact that the basic Skyrme model is consistent at low momentum and the assumed homogeneity of the fluctuation. This is taken into account by taking for $\Omega ^{-1}(x,y,t)$ the cut-off kernel \begin{equation} \Omega ^{-1} (x,y,t) = \frac{1}{2\pi \Omega _{0}} \int _{-K}^{K} e^{ik(x-y)} \: \left( k^{2} + a^{2} \right) \: dk. \label{e:kernal} \end{equation} Then the non-constant contribution of $\Omega ^{-1}(x,x,t)$ is \begin{equation} -\frac{1}{8} \Omega ^{-1} (x,x,t) = -\frac{K}{8\pi \Omega_{0}} a^{2}. \label{e:contrib} \end{equation} In a similar manner we obtain \begin{eqnarray} \frac{1}{2} \frac{\partial ^{2}}{\partial x^{2}} \Omega (x,y,t) |_{y=x} & = & \lim _{K \to \infty} -\frac{1}{4\pi} \Omega_{0} \int _{-K}^{K} \frac{k^{2}} {k^{2}+a^{2}} \: dk \nonumber \\ & = & \lim _{K \to \infty} \left[ -\frac{\Omega_{0} K}{2\pi} + \frac{1}{4\pi} a(t) \Omega_{0} \int _{-\infty}^{\infty} \frac{dk}{k^{2}+1} \right] . \label{e:limit2} \end{eqnarray} The first term in this expression is infinite, but a constant. Thus it does not contribute to the equations of motion. We therefore take just the second term in the effective Lagrangian. Finally the parameter $\Omega_0$ measures the size of the fluctuations. In principle, other choices of the parameters may lead to different regularizations. However, as discussed below, the basic qualitative picture described in this work is not changed by these alternative regularizations, provided the momentum is low. For the case of higher momentum, procedures similar to the ones discussed in \cite{eboli2,cooper} can be used. With the above assumptions, the effective Lagrangian takes the form \begin{eqnarray} L &=& \int_{0}^{T} \int_{-L}^{L} \left\{ \pi \frac{\partial \varphi}{\partial t} - \left( \frac{1}{2} \left( \frac{\partial \varphi}{\partial x}\right)^{2} + \left[ 1 - \left( 1-\frac{\Omega_{0}}{2\pi a} c_{2}\right) \cos \varphi \right] \right) \right. \nonumber\\ & & \mbox{} \left. + \frac{1}{2\pi} \left( - 2 \frac{\dot{a} b}{a^{4}}\Omega_{0} c_{1} -2\frac{b^{2}}{a^{5}} \Omega_{0} c_{1} - \frac{K}{4\Omega_{0}} a^{2} + \frac{1}{2} \Omega_{0} c_{2} a \right)\right\} \: dx dt, \label{e:efflang} \end{eqnarray} where the constants $c_{1}$ and $c_{2}$ are given by \begin{equation} c_{1}= \int_{-\infty}^{\infty} \frac{dk}{(1+k^{2})^{3}}=\frac{\pi}{2}, \qquad c_{2}= \int_{-\infty}^{\infty} \frac{dk}{(1+k^{2})}= \pi. \label{e:const} \end{equation} It is now convenient to change variables and define $q=\frac{1}{a^{3}}$ and $b=p$ in order to obtain the Lagrangian (\ref{e:efflang}) in the form \begin{eqnarray} L &=& \int_{0}^{T} \left\{ \left( \frac{2 c_{1} \Omega_{0}}{3} \dot{q} p - 2c_{1}\Omega_{0} p^{2} q^{-\frac{5}{3}} - \frac{K}{4\Omega_{0}} q^{-2/3} + \frac{1}{2} \Omega_{0} \frac{c_{2}}{c_{1}} q^{-1/3} \right) \: \frac{L}{\pi} dt \right. \nonumber \\ & & \mbox{} + \left. \int_{-L}^{L} \left[ \pi \frac{\partial \varphi} {\partial t}- \frac{1}{2} \pi^{2} - \frac{1}{2} \left( \frac{\partial \varphi} {\partial x}\right)^{2} - [1-(1- \frac{\Omega_{0}}{2\pi} \pi q^{\frac{1}{3}}) \cos \varphi] \right] \: dx \right\} \: dt. \label{e:chlang} \end{eqnarray} The equations of motion are obtained by varying the Lagrangian (\ref{e:chlang}) with respect to the parameters $\pi$, $\varphi$, $p$ and $q$. These variational equations will consist of a partial differential equation for the field coupled to ordinary differential equations for the fluctuations. For the field the variational equation is \begin{equation} \frac{\partial^{2} \varphi}{\partial t^{2}} - \frac{\partial^{2} \varphi}{\partial x^{2}} + \left( 1 - \frac{\Omega_{0}}{2} q^{\frac{1}{3}}\right) \sin {\varphi} = 0. \label{e:psineg} \end{equation} The equations for $p$ and $q$ are derived from the Lagrangian \begin{equation} {\cal L} = \frac{2c_{1}L\Omega_{0}}{3\pi} \int _{0}^{T} \left[ \dot{q} p - 3 H(p,q) \right] \: dt, \label{e:pqlang} \end{equation} where the Hamiltonian for the fluctuations is given by \begin{equation} H(p,q) = p^{2} q^{-5/3} + q^{1/3} \frac{1}{2L} \int _{-L}^{L} \cos \varphi \: dx - \frac{1}{8} q^{-1/3} + \frac{K}{8c_{1} \Omega^{2}_{0}} q^{-2/3}. \label{e:hamfl} \end{equation} The equations of motion for $p$ and $q$ are then given by Hamilton's equations as \begin{eqnarray} \dot{q} & = & \frac{\partial H}{\partial p} = 6p q^{-5/3} \label{e:peqn} \\ \dot{p} & = & -\frac{\partial H}{\partial q} = 5 p^{2} q^{-8/3} + \frac{2}{3} \beta q^{-5/3} - \frac{1}{24} q^{-4/3} - \frac{q^{-2/3}}{2L} \int_{-L}^{L} \cos \varphi \: dx, \label{e:qeqn} \end{eqnarray} where \begin{equation} \beta = \frac{3K}{8c_{1} \Omega^{2}_{0}}. \label{e:beta} \end{equation} Since $K$ is assumed to be small, we will take $\beta = 2.5$ in the numerical calculations of the next section. It must be noted that for initial conditions which have $\varphi \to 2 \pi$ as $x \to -\infty$ and $\varphi \to 0$ as $x \to \infty$, or vice versa, \begin{equation} \Lambda:=\lim_{L \to \infty} \frac{1}{2L} \int _{-L}^{L} \cos \varphi \: dx = 1. \label{e:lim1} \end{equation} It is therefore apparent that the Equations (\ref{e:qeqn}) for the fluctuations decouple from the Equation (\ref{e:psineg}) for the field. It is interesting at this point to compare our equations (\ref{e:psineg}), (\ref{e:peqn}) and (\ref{e:qeqn}) with the corresponding equations (4.5) of Ref. \cite{cooper}. Observe that in that reference the equation for the field takes the form \begin{equation} \varphi_{tt}-\varphi_{xx}+\frac{\alpha_0}{\beta} e^{-\frac{\beta^2}{2}G(x,x,t)}\sin \beta\varphi=0. \label{e:mam} \end{equation} Taking $\beta=1$ and assuming $G(x,x,t)\ll 1$, which amounts to choosing the initial conditions in the form of small fluctuations, we have \begin{equation} e^{-\frac{1}{2}G(x,x,t)}\approx 1-\frac{1}{2}G(x,x,t). \end{equation} Clearly, substituting this last expression into (\ref{e:mam}) recovers our equation (\ref{e:psineg}), so the respective field equations agree for the above mentioned initial conditions and value of $\beta$. Now, as for the second equation (4.5) in \cite{cooper} note that this is an infinite system of partial differential equations for the operator valued function $\hat{\Psi}$, while in our formulation, because of the assumption of spatial homogeneity and the functional form chosen for the trial Green function, the system simplifies to a problem of ordinary differential equations. Another difference between our approach and that followed in \cite{cooper} is that the equations proposed there in order to arrive at numerical solutions are integrodifferential equations, as opposed to the simple variational approximation we propose for obtaining solutions. It must be also remarked that our assumed homogeneity of the Green function is consistent with the low momentum limit we have chosen. In this low momentum limit, the fluctuations do not resolve the fine scale of the field and, to leading order, the configuration is a homogeneous background for the fluctuation. To conclude this section we also consider important to stress the fact that in previous published work the interest has been on static solutions allowing for arbitrary momentum of arbitrarily large fluctuations. This leads to a different renormalized version of the gap equation \cite{eboli1,eboli2,cooper} and, for large coupling $\beta$, to a loss of stability (phase transition). Our approximation does not capture this region, since we have assumed from the onset small fluctuations and small momentum. However it must be noted that our procedure could be extended to handle large momenta by choosing different trial functions for $G$ and $\Sigma$, similar to the ones used in \cite{eboli2}. Also, due to the special form of the potential in the equations, the Gaussian integral may be evaluated to a better degree of approximation, thus allowing to include fluctuations of a larger amplitude. This program is currently under investigation and will be reported subsequently. In the following section we undertake a detailed study of the dyamics described by the equations (\ref{e:psineg}), (\ref{e:peqn}) and (\ref{e:qeqn}). \section{SOLUTIONS} The system of the Sine-Gordon equation (\ref{e:psineg}) and the equations (\ref{e:qeqn}) and (\ref{e:peqn}) for $q$ and $p$ describe the coupling of the field to the fluctuations and the corresponding feedback. Note that the fluctuations have been assumed to be small. However, they are allowed to feed back onto the basic field configuration. We shall now use these equations to describe in a nonperturbative manner the nonlinear evolution of some special field configurations. \subsection{Quantum Stability of the Single Soliton} We begin by studying the stability of the Soliton solution (\ref{e:sgsol}) under a class of initial values for $p$ and $q$. Note that a small value of $q$ represents a small variance. Stability is then assured in the model by the fact that $q$ remains small and that the field maintains its identity as a localized structure. Numerical integrations of the Sine-Gordon equation (\ref{e:psineg}) and the Equations (\ref{e:qeqn}) and (\ref{e:peqn}) for $p$ and $q$ have been performed for a wide range of initial conditions and typical behaviors are shown in Figures \ref{f:sol} and \ref{f:solmax}. In Figure \ref{f:sol} a numerical solution for $\Phi = \varphi _{x}$ for the Skyrmion is shown and in Figure \ref{f:solmax} the behavior of $a$, the maximum of $\Phi$, is shown. It can be seen that the fluctuations of the Skyrmion produce radiation, but that the field eventually stabilizes. This can be clearly seen from the maximum behavior shown in Figure \ref{f:solmax}. The stabilization onto a modulated small oscillation of the Skyrmion amplitude can be clearly seen. These results exhibit the strong stability of the Skyrmion with respect to fluctuations. It is possible to understand this behavior by making use of the modulation theory given in \cite{minzoni} and \cite{whitham} by means of the following argument. If the scales for $p$ and $q$ are slow, we may take as an approximate solution \begin{equation} \varphi = \varphi \left( (1 - \frac{\Omega_{0}}{2} q^{\frac{1}{3}})^{\frac{1}{2}} x \right), \label{e:appsoln} \end{equation} which satisfies \begin{equation} \cos \varphi = 1 - \frac{1}{2} \dot{\varphi}^{2}. \label{e:cosphi} \end{equation} Hence as $L \rightarrow \infty$ we have from (\ref{e:lim1}) that $\Lambda=1$. With this, the equations for $p$ and $q$ are derivable from the Hamiltonian \begin{equation} H(p,q)=3\left( p^{2} q^{-\frac{5}{3}} + q^{1/3} - \frac{1}{8} q^{-1/3} + \frac{K}{4\Omega^{2}_{0}} \: q^{-2/3} \right). \label{e:appham} \end{equation} Thus the orbits in the $(p,q)$ plane are just the level lines of \begin{equation} H(p,q) = E. \label{e:orbits} \end{equation} The orbits of the $(p,q)$ system are then given by \begin{eqnarray} p^{2} & = & q^{5/3} \left[ E - \left( \frac{K}{4\Omega^{2}_{0}} \: q^{-2/3} + q^{1/3} - \frac{1}{8} q^{-1/3}\right) \right] \nonumber \\ & = & q^{5/3} \left[ E - V(q) \right] . \label{e:pqham} \end{eqnarray} The potential $V(q)$ has a minimum which gives an oscillatory solution for $p$ and $q$, so that the width of the Skyrmion and thus the amplitude of $\Phi = \varphi _{x}$ oscillate in time. The numerical solution shown in Figure 1 shows that the radiation, which is not taken into account in this approximation, stabilizes the oscillations onto a limit cycle. This strongly nonlinear mechanism accounts for the stability of the Skyrmion. In fact it is the feedback of the field on the fluctuations which produces the term $q^{1/3}$ in $V(q)$ and it is this term which stabilizes the motion. The potential $V(q)$ has a maximum for small $q$. For energies $E$ larger than this maximum, the fluctuations $q$ increase and the field structure is destroyed. However the value of $q$ for this to occur is too small to be consistent with the coherent state approximation. The model is therefore self-consistent and provides an explanation of how nonlinear interactions are responsible for the quantum stability of the field. The approximate solution above does not take into account the radiation produced by the oscillating Skyrmion and so this approximate solution will not give the baryon settling onto a limit cycle solution. To take account of the radiation the ideas of Smyth and Worthy \cite{annette} can be used. In this work the effect of shed dispersive radiation on the evolution of a single pulse for the Sine-Gordon equation was treated. To take account of the radiation we proceed as in \cite{annette}, indicating only the main differences from this work. The Lagrangian density for the Sine-Gordon equation is \begin{equation} L = \frac{1}{2} \varphi_{t}^{2} - \frac{1}{2} \varphi _{x}^{2} - \left( 1 - \frac{\Omega_{0}}{2} q^{1/3} \right) \left( 1 - \cos \varphi \right) . \label{e:sglang} \end{equation} To obtain an approximate solution of the Sine-Gordon equation, the trial function \begin{equation} \varphi = -4 \mathop{\rm arctan}\nolimits e^{-x/w(t)}, \label{e:trial2} \end{equation} which is a soliton-like pulse with varying width $w(t)$, is substituted into the averaged Lagrangian \begin{eqnarray} \bar{L} & = & \int_{-\infty}^{\infty} L \: dx \nonumber \\ & = & \frac{\pi^{2}}{3} \frac{w'^{2}}{w} - \frac{4}{w} - 4\left( 1 - \frac{\Omega_{0}}{2} q^{1/3} \right) w , \label{e:annlang} \end{eqnarray} as in \cite{annette}. In this approximation the Hamiltonian for $p$ and $q$ again does not change due to (\ref{e:lim1}) The effect of the radiation shed by the evolving soliton is determined by finding an appropriate solution of the linearized Sine-Gordon equation \cite{annette}. The effect of this radiation is then to modify the Euler-Lagrange equation for $w(t)$. It is noted from the numerical solution of Figure \ref{f:sol} that the radiation $\tilde{\varphi}$ is of small amplitude compared with the soliton. Therefore following \cite{annette} we consider the linearized Sine-Gordon equation \begin{equation} \frac{\partial ^{2} \tilde{\varphi}}{\partial t^{2}} - \frac{\partial ^{2} \tilde{\varphi}}{\partial x^{2}} + \left( 1 - \frac{\Omega_{0}}{2} q^{1/3}(t) \right) \tilde{\varphi} = 0 \label{e:linsg} \end{equation} for the radiation $\tilde{\varphi}$. This equation is solved together with appropriate source conditions at the pulse at $x=0$. Since $\Omega_{0} \ll 1$, \begin{equation} \frac{d}{dt} \left( 1 - \frac{\Omega_{0}}{2} q^{1/3}(t) \right) \ll 1. \label{e:ll1} \end{equation} It is then possible to obtain an expression for the radiation by making the adiabatic approximation that $1 - \frac{\Omega_{0}}{2} q^{1/3}$ is constant to leading order. The effect of the radiation can then be found from the expression of \cite{annette} by a suitable re-scaling. In this manner we find that the equations governing the evolution of the soliton, including the effect of radiation, are \begin{eqnarray} \frac{2\pi^{2}}{3w} \frac{d^{2}w}{dt^{2}} - \frac{\pi^{2}}{3w^{2}} \left( \frac{dw}{dt} \right) ^{2} - \frac{4}{w^{2}} & & \nonumber \\ \mbox{} + 4 \left( 1 - \frac{\Omega_{0}}{2} q^{1/3} \right) & = & \frac{1}{\sqrt{\lambda}} \left[ -\frac{2\pi^{2}}{3w\sqrt{\lambda} \: t} \frac{dw}{dt} \right. \nonumber \\ & & \mbox{} + \left. \frac{2\pi^{2}}{3\sqrt{wt}} \int_{0}^{t} J_{1}(\sqrt{\lambda}(t-\tau)) \frac{w'(\tau)}{\sqrt{\tau w(\tau)}} \: d\tau \right] \nonumber \\ \frac{dq}{dt} & = & 6 p q^{-5/3} \label{e:radeqn} \\ \frac{dp}{dt} & = & 5 p^{2} q^{-8/3} + \frac{2}{3} \beta q^{-5/3} - q^{-2/3} + \frac{1}{9} q^{-4/3}, \nonumber \end{eqnarray} where \begin{equation} \lambda = 1 - \frac{\Omega_{0}}{2} q^{1/3}. \label{e:lam} \end{equation} These equations were integrated numerically. Comparisons between solutions of these equations and the full numerical solution of the Sine-Gordon equation for the amplitude $a$ of $\Phi = \varphi _{x}$ and $q(t)$ for the fluctuations are shown in Figure \ref{f:ampcomp}. It can be seen that the amplitude agreement shown in Figure \ref{f:ampcomp} is good considering the assumptions that were made to incorporate the radiation loss in the approximate equations. It can be seen that the approximate equations provide a suitable approximate solution for the full field behaviour using a finite dimensional approximation which includes radiation. It is noted that, since $q(t)$ is periodic, the Sine-Gordon Equation (\ref{e:psineg}) is subject to a parametric excitation. However the nonlinearity and radiation loss provide the necessary damping to enable a limit cycle to be achieved. \subsection{The collision of a wave with a static soliton} As a final example we consider the scattering of a wave packet representing a pion with momentum $k$ with a static soliton, representing a baryon originally at rest. The problem is solved by numerically integrating the Sine-Gordon equation (\ref{e:psineg}) using the initial condition \begin{eqnarray} \varphi(x) & = & -4\mathop{\rm arctan}\nolimits e^{-x}+f(x) \label{e:icf} \\ \frac{\partial\varphi}{\partial t} & = & g(x), \label{e:icv} \end{eqnarray} where the functions $f$ and $g$ are given by \begin{eqnarray} f(x) & = & a\sin k(x+x_0),\quad |x+x_0|\leq\delta \label{e:f} \\ g(x) & = & -a\sqrt{k^2+1}\cos k(x+x_0),\quad |x+x_0|\leq\delta . \label{e:g} \end{eqnarray} This initial condition represents an incoming meson with momentum $k$ impinging on a nucleon located at $x=0$. A numerical solution for the scattering of the pion wavepacket can be seen in Figures \ref{f:solrad} to \ref{f:raddet}. The initial condition (at $t=0$) is shown by the solid line in Figure \ref{f:solrad}. In this Figure a reflected wave packet, a reorganized field configuration and a new packet shed by the baryon as a result of the interaction can be seen. In Figures \ref{f:radmax} and \ref{f:raddet} the complicated evolution of the baryon amplitude is displayed. This amplitude behavior is due to the interaction of the baryon with the packet. The scattering then involves a reorganization of the field, which is not taken into account when the particles are taken to be point particles. The description of the interaction of the baryon with radiation using a multi-phase solution of the Sine-Gordon equation is under investigation at present. \subsection{Collision of two solitons in the presence of a fluctuation} Since the classical field equation is completely integrable, solitons interact elastically and do not change configuration. The effect of quantum fluctuations on the collisions of solitons and this clean interaction will now be studied. Figure \ref{f:two} shows the collision of two Solitons with equal and opposite velocity. The initial condition used was \begin{equation} \varphi = 2 \pi - 4 \mathop{\rm arctan}\nolimits e^{-(x+x_{0}-vt)/\sqrt{1-v^{2}}} - 4 \mathop{\rm arctan}\nolimits e^{-(x-x_{0}+vt)/\sqrt{1-v^{2}}} \label{e:twosol} \end{equation} as $t \to -\infty$. Since there is no classical solution with twice the baryon number and zero velocity, the effect of the quantum fluctuations is just to slightly modify the classical interaction. The Solitons again settle down to a limit cycle for which the parametric resonance is balanced by the radiation damping. Figure \ref{f:anti} shows the collision of a Soliton and an anti-soliton. The initial condition is \begin{equation} \varphi = -4 \mathop{\rm arctan}\nolimits \left[ \frac{\alpha}{\sqrt{\alpha ^{2} -1}} \frac{\sinh \sqrt{\alpha^{2}-1} \: t}{\cosh \alpha x} \right]. \label{e:anti} \end{equation} Again this interaction does not produce disintegration, just a modification of the classical interaction. Finally the susceptibility to disintegration of the breather-type configuration \begin{equation} \varphi = -4 \mathop{\rm arctan}\nolimits \left[ \frac{\alpha}{\cosh \alpha x} \frac{\sin \sqrt{1-\alpha^{2}} \: t}{\sqrt{1-\alpha^{2}}} \right] \label{e:breath} \end{equation} with frequency $\sqrt{1-\alpha^{2}}$ is studied. From the numerical solution shown in Figure \ref{f:breath} it can be seen that the breather is stable with respect to quantum fluctuations. The solutions displayed in Figure 4 show that the reduction of the Skyrme model to the Sine-Gordon equation is too severe for treating collisions. In order to obtain non-trivial collision and fusion processes, such as those possible for the nonlinear Schr\"odinger equation, reductions of the Skyrme model which retain more internal degrees of freedom must be derived. \section{CONCLUSIONS AND SUGGESTIONS FOR FURTHER RESEARCH} We have formulated the quantum field problem for the Sine-Gordon equation which is related to the (dimensionally reduced) Skyrme model. Using the coherent state approximation for the solution of the functional Schr\"odinger equation, we obtain a solution of the partial differential equation for the (quantum corrected, semi-classical) field, which is coupled to ordinary differential equations for the fluctuations. Other quantizations for nonlinear fields keep only finitely many degrees of freedom (minisuperspace approximation), which are then quantized in a canonical way. The first problem considered in the present work was the stability with respect to quantum fluctuations of a Soliton. Both numerical and asymptotic solutions were considered. It was found that the nonlinear saturation of the field equation together with the loss of radiation balanced the parametric excitation of the fluctuations. The fluctuations in turn were controlled by the shape of the field. The good agreement found between numerical and asymptotic solutions suggests that finite dimensional approximations to the dynamics of the Skyrmion model, such as those used in \cite{gisiger}, are also good approximations to the full dynamics of more complicated problems, such as those treated in \cite{gisiger}. The scattering of a wave by a static soliton was also studied. The numerical results obtained show well defined waves and a Skyrmion after collision, which suggests the possibility of using multi-phase solutions, such as those of \cite{15}, to understand this scattering process. Finally several collision processes were studied. It was found that the reduced Skyrme model cannot account for the collision and fusion of baryons. Therefore the study of the fusion of Skyrmions into a toroidal configuration requires a uniform solution which interpolates between the torus and the individual Skyrmions. The possibility of using the solutions given in \cite{13,14} is currently under study. It must be noted that more sophisticated numerical formulations such as the ones proposed in \cite{cooper} must produce, in the limit of low momentum, solutions comparable to our results. To conclude, we note that the techniques described in this work can be applied to the study of low dimensional black holes. Indeed, an old observation that Sine-Gordon theory and 2-dimensional spaces of constant curvature are very closely related has recently found an interesting application to gravity in $1+1$ dimensions. More precisely, Gegenberg and Kunstatter \cite{gegen} have noticed that when a two dimensional Lorenzian metric is parameterized as \begin{equation} ds^2 = -\sin^2 (u/2) dt^2 + \cos^2 (u/2) dx^2, \label{e:metric} \end{equation} then the condition of constant curvature is equivalent to the condition that $u$ satisfies the {\it Euclidean} Sine-Gordon equation. On the other hand, the so-called Jackiw-Teitelboim theory in two dimensions \begin{equation} I = \int \phi(R - \Lambda )\sqrt{-g} \: dt dx. \label{e:jack} \end{equation} has as solutions space-times of constant curvature $R=\Lambda$. Furthermore, the one-soliton solution of the Sine-Gordon equation has been found to represent (a patch of) a black hole solution of the (Jackiw-Teitelboim) theory \cite{gegen}. That a constant curvature space-time can be interpreted as a black hole is not unique to two dimensions. The $1+1$ Jackiw-Teitelboim black hole can indeed be interpreted as a dimensionally reduced BTZ (non-rotating) black hole and many of its properties (including thermodynamics) have been studied \cite{lemos}. To perform an analysis similar to the one presented in the present work for the Euclidean Sine-Gordon equation is cumbersome, since the equation is now elliptic and does not accept a well-posed initial value formulation. However it is possible to work in the framework of a well-posed problem if one chooses a different parametrization for the two dimensional space-time as follows \begin{equation} ds^2 = -\sinh^2 (u/2) dt^2 + \cosh^2 (u/2) dx^2. \label{e:dssinh} \end{equation} In this case, the constant curvature condition reduces to the {\it Lorenzian} Sinh-Gordon equation. It is then possible to analyze the quantum stability of a black hole solution using the functional methods presented in this article. This work will be reported elsewhere. As a final remark we point out that the quantum equations for a classical field obtained using the functional Schr\"odinger equation and the coherent state approximation will always have the same structure. Namely the classical equations for the field with renormalized (fluctuating) parameters and equations for the (parameters of the) fluctuations which are non-local in the fields will always be obtained. \section{ACKNOWLEDGEMENTS} This work was supported in part by UNAM DGAPA Project No.\ IN 106097. A.C. was also supported by Conacyt Proyect No.\ I25655-E.
\section{Introduction} \labell{sec:intro} Abstract moment maps arise as a generalization of genuine moment maps on symplectic manifolds. The essence of their definition is that the symplectic structure is discarded but the relation between the mapping and the action is kept intact. To be precise, an abstract moment map on a $G$-manifold $M$ is an equivariant mapping $\Psi\colon M\to{\mathfrak g}^*$ satisfying the following constancy condition: for every Lie algebra element $\xi\in{\mathfrak g}$, the component $\left<\Psi,\xi\right>$ is locally constant on the set $\{\xi_M=0\}$ where the action generating vector field $\xi_M$ vanishes. Moment maps on symplectic manifolds are among examples of abstract moment maps. In general, however, an abstract moment map is not associated with a symplectic form or even a closed two-form. An abstract moment map is an additional structure on a $G$-manifold, and a given $G$-manifold admits many abstract moment maps. Thus $G$-manifolds equipped with abstract moment maps occupy an intermediate place between pure $G$-manifolds and symplectic manifolds with Hamiltonian $G$-actions. The goal of this paper is to find a relationship between $G$-manifolds, $G$-manifolds equipped with abstract moment maps, and Hamiltonian $G$-spaces. This is done by using a new notion, the notion of an \emph{assignment} comprising certain combinatorial data extracted from an abstract moment map. An assignment should be thought of as a combinatorial counterpart of an abstract moment map. For a torus action, an assignment is a function from the set of orbit type strata to the dual spaces to the Lie algebras of the stabilizers. (Thus a manifold with a finite orbit type stratification has only a finite--dimensional space of assignments.) First we address the existence and uniqueness question for abstract moment maps with a fixed assignment. We prove that, for a torus action, every assignment is associated with an abstract moment map. Furthermore, two abstract moment maps give rise to the same assignment if and only if they differ, roughly speaking, by a Hamiltonian moment map arising from an exact two-form. For some problems concerning non-compact manifolds, it is important to consider abstract moment maps which are proper. (Non-compact $G$-manifolds with proper abstract moment maps share some of the appealing properties of compact $G$-manifolds. See \cite{Ka,GGK:lerman,GGK:book}.) We show that for a given assignment, an abstract moment map can be chosen proper if the assignment is proper or, to be more precise, ``polarized"; see Section \ref{subsec:polar}. Then we use the notion of an assignment to answer the question whether a given abstract map is associated with a two-form. Moment maps associated with true symplectic forms must additionally satisfy some non-degeneracy requirements. These requirements are analyzed in \cite{GGK:book}. On the other hand, moment maps on Poisson manifolds (see, e.g., \cite{wein-lect}) are not in general abstract moment maps because they do not have to be locally constant on the fixed point set. The paper is organized as follows. In Section \ref{sec:def-examples} we recall the definition of abstract moment maps and illustrate it by a number of examples. In Section \ref{sec:existence} we give a necessary and sufficient condition (in terms of assignments) for a $G$-manifold to admit a (polarized) abstract moment map. In Section \ref{sec:exact} we show that two abstract moment maps with the same assignment differ by one which is associated with an exact two-form (or, to be more precise, with a one-form). We call such abstract moment maps \emph{exact}. Section \ref{sec:forms} is devoted to the question of which abstract moment maps are Hamiltonian, i.e., associated with closed two-forms. We show that every abstract moment map is locally Hamiltonian. Globally, there is an obstruction, which is stated again in terms of assignments. In Section \ref{sec:proof} we prove a technical theorem on which the results of Sections \ref{sec:exact} and \ref{sec:forms} heavily rely: an abstract moment map on a linear representation is exact if and only if it vanishes at the origin. Finally, we show that the space of assignments and some of its generalizations fit as the zeroth cohomology in a series of certain cohomology spaces associated with a $G$-manifold. This cohomology is introduced and studied in Section \ref{sec:cohomology}. Abstract moment maps were introduced in \cite{Ka} to study geometric equivariant $G$-cobordisms and to state and prove the cobordism linearization theorem. This theorem, whose earliest version was given in \cite{GGK1} (see also \cite{GZ} for important related work), asserts that under certain hypotheses a manifold with a torus action is equivariantly cobordant to the disjoint union of the linear isotropy representations at the fixed points.\footnote{ A number of applications of the linearization theorem are outlined in \cite{GGK1} and \cite{Ka}. See \cite{MW1} and \cite{MW2} for some new applications, and see \cite{GGK:book} and \cite{GGK:lerman} for further developments. } One of the main conceptual difficulties in the formulation of this theorem is to find a notion of non-compact cobordism which would not render every compact manifold cobordant to the empty set. (This is the central problem arising when non-compact manifolds are introduced in a cobordism theory: all compact manifolds may become cobordant to each other and so to zero.) The notion of an abstract moment map provides a solution to this problem: two $G$-manifolds equipped with proper abstract moment maps are cobordant if there exists a $G$-cobordism between them and a proper abstract moment map on it extending those on the boundary. This definition leads to a non-trivial cobordism theory in which the theory of compact (geometric) $G$-cobordisms is embedded.\footnote{ Strictly speaking, this is true for geometric stable complex $G$-cobordisms. See \cite{GGK:lerman} and references therein. } The non-compact theory appears to be in some sense simpler than the compact one. The reason is that the non-compact theory has a well-understood set of generators and probably fewer relations than the compact one. (See \cite{Ka} and \cite{GGK:book}.) We feel, however, that, as some examples in Sections \ref{sec:def-examples} and \ref{sec:cohomology} indicate, abstract moment maps and assignments may have uses beyond those connected with geometric equivariant $G$-cobordisms. \subsection*{Notation and conventions} Throughout this paper, $M$ is a $G$-manifold ($C^{\infty}$-smooth), with $G$ being a torus, except in some rare cases where $G$ is allowed to be a more general compact Lie group. As usual, ${\mathfrak g}$ denotes the Lie algebra of $G$ and $\xi_M$ is the vector field induced by the action of $\xi\in{\mathfrak g}$ on $M$. The stabilizer of $x\in M$ is denoted $G_x$ and the fixed point set of $G$ on $M$ by $M^G$. All ordinary and equivariant cohomology groups are assumed to have real coefficients unless specified otherwise. We consider only abstract moment maps for torus actions. Many (but not all) of our results should extend to proper actions of other Lie groups. \subsection*{Acknowledgment} The authors wish to thank Phil Bradley, Emmanuel Farjoun, Debra Lewis, Assaf Libman, David Metzler, Avishay Vaaknin, and Yuli Rudyak for fruitful discussions and comments. \section{Abstract moment maps} \labell{sec:def-examples} Let us first recall some standard facts about smooth group actions. Let a Lie group $G$ act smoothly on a manifold $M$. Assume that $G$ is compact or, more generally, that the action is proper (that is, that the map $(a,m) \mapsto (a \cdot m , m)$ from $G \times M$ to $M \times M$ is proper). Each Lie algebra element $\xi \in {\mathfrak g}$ gives rise to a vector field $\xi_M$ on $M$. The \emph{stabilizer} of a point $m \in M$ is the group $\{a \in G \ | \ a \cdot m = m \}$. The Lie algebra of this group is equal to $\{ \xi \in {\mathfrak g} \ | \ \xi_M|_m = 0 \}$ and is called the \emph{infinitesimal stabilizer} of $m$. For each subgroup $H \subseteq G$, the connected components of the set of points whose stabilizer is conjugate to $H$ (hence equal to $H$, if $G$ is Abelian) are smooth sub-manifolds of $M$. These connected components are the \emph{orbit type strata} of $M$. They are partially ordered: $X \preceq Y$ if and only if the stratum $X$ is contained in the closure of the stratum $Y$. Similarly, the connected components of the sets of points whose infinite stabilizers are conjugate to given sub-algebras ${\mathfrak h} \subseteq {\mathfrak g}$ (hence equal to ${\mathfrak h}$, if $G$ is Abelian) form the \emph{infinitesimal obit type strata} of $M$. These are, too, partially ordered by inclusions of closures of strata. The infinitesimal orbit type stratification is more coarse than the orbit type stratification, because two points $x$ and $y$ can have different stabilizers but the same infinitesimal stabilizer. In what follows we will mainly work with the infinitesimal orbit type stratification because this stratification is more suitable for the goals of this paper than the orbit type stratification. Next, let us recall the definition of abstract moment maps, as given in \cite{Ka}. For a map $\Psi \colon M \to {\mathfrak g}^*$, we denote by $\Psi^{\mathfrak h}$ or $\Psi^H$ the composition of $\Psi$ with the natural projection ${\mathfrak g}^* \to {\mathfrak h}^*$, where ${\mathfrak h}$ is the Lie algebra of $H$. Similarly, for any Lie algebra element $\xi \in {\mathfrak g}$, we denote by $\Psi^\xi \colon M \to {\mathbb R}$ the $\xi$th component of $\Psi$, i.e., $\Psi^\xi=\left<\Psi,\xi\right>$. \begin{Definition} \labell{def:moment} An \emph{abstract moment map} on $M$ is a smooth map $ \Psi \colon M \to {\mathfrak g}^* $ with the following properties: \begin{enumerate} \item $\Psi$ is $G$-equivariant, and \item for any subgroup $H$ of $G$, the map $\Psi^H \colon M \to {\mathfrak h}^* $ is locally constant on the submanifold $M^H$ of points fixed by $H$. \end{enumerate} \end{Definition} \begin{Remark} For the second requirement to hold, it is enough to assume that for any Lie algebra element $\xi \in {\mathfrak g}$, the function $\Psi^\xi$ is locally constant on the set of zeros of the corresponding vector field $\xi_M$. If $G$ is compact, it is enough to demand the requirement for circle subgroups of $G$. \end{Remark} In this paper we mainly consider the case where $G$ is a torus and we often focus on abstract moment maps that are proper. \begin{Example} \labell{ex:zero} The constant function zero is an abstract moment map. It is proper if and only if $M$ is compact. \end{Example} \begin{Example} If the fixed point set $M^G$ has a non-compact component, $M$ does not admit a proper abstract moment map. \end{Example} \begin{Example} Let $G$ be the circle group, and let us identify ${\mathfrak g}^*$ with ${\mathbb R}$. Then an abstract moment map is a real valued invariant function that is constant on each connected component of the fixed point set. In particular, if the set of fixed points is discrete, any invariant function is an abstract moment map. \end{Example} \begin{Example} \labell{exam:presymplectic} Recall that a Hamiltonian $G$-space is a triple $(M,\omega,\Psi)$, where $M$ is a $G$-manifold, $\omega$ is a closed invariant two-form (which in some contexts -- not here -- is required to be symplectic), and $\Psi$ is a moment map, i.e., a $G$-equivariant function $\Psi \colon M \to {\mathfrak g}^*$, such that Hamilton's equation, \begin{equation} \labell{Hamilton} \iota(\xi_M) \omega = - d\Psi^\xi \quad \text{for all $\xi \in {\mathfrak g}$}, \end{equation} holds. Then $\psi$ is an abstract moment map. \end{Example} If equation \eqref{Hamilton} holds, we say that $\omega$ is compatible with $\Psi$, or that $\Psi$ is associated with $\omega$. An abstract moment map associated with some two-form will be called a \emph{Hamiltonian} moment map. \begin{Example} \labell{ex:one-form} Let a Lie group $G$ act on a manifold $M$ and let $\mu$ be any invariant one-form. Then the function $\Psi \colon M \to {\mathfrak g}^*$ defined by \begin{equation} \labell{moment one form} \Psi^\xi = \mu(\xi_M) \end{equation} is an abstract moment map. Moreover, for each $H \subset G$, the function $\Psi^H$ vanishes on $M^H$. \end{Example} \comment{\ In the book, expand the following paragraph:} An abstract moment map that arises by Equation \eqref{moment one form} is called \emph{exact}. A compatible two-form is then given by $\omega = d\mu$. Many of ``classical'' moment maps, e.g., the canonical moment map on the cotangent bundle, are exact. Also, in the pre-quantization of a Hamiltonian action, the pullback of the moment map to the pre-quantum circle bundle is an exact abstract moment map. The advantage of working with exact moment maps over Hamiltonian ones is that if $\Psi_0$ and $\Psi_1$ are exact moment maps then so is $(1 - \rho) \Psi_0 + \rho \Psi_1$ for any smooth function $\rho$. \begin{Remark} Recall that for any Lie group $G$ an equivariant differential two-form on a $G$-manifold $M$ is a formal sum \begin{equation} \labell{omegaPsi} \omega + \Psi, \end{equation} where $\omega$ is an invariant two-form on $M$ and $\Psi$ is a smooth equivariant function from $M$ to ${\mathfrak g}^*$. The equivariant form \eqref{omegaPsi} is said to be equivariantly closed if and only if it satisfies \eqref{Hamilton}; it is said to be equivariantly exact if and only if there exists an invariant one-form $\mu$ such that $\omega = d\mu$ and $\Psi^\xi = \mu(\xi_M)$ for all $\xi \in {\mathfrak g}$. The second equivariant cohomology, denoted $H^2_G(M)$, is the quotient of the space of equivariantly closed equivariant two-forms by the subspace of those that are equivariantly exact. \end{Remark} \begin{Example} \labell{pull-back} The pull-back of an abstract moment map is an abstract moment map. More precisely, let $f\colon N\to M$ be an equivariant map of $G$-manifolds and let $\Psi$ be an abstract moment map on $M$. Then $f^*\Psi=\Psi\circ f$ is an abstract moment map on $N$; the map $f^*\Psi$ is proper, provided that $f$ and $\Psi$ are proper. For instance, following \cite{SLM} and \cite{Le}, consider a $G$-manifold $Q$ and denote by $J\colon T^*Q\to {\mathfrak g}^*$ the canonical moment map: $J^\xi(\mu)=\mu(\xi_Q(x))$, where $\mu\in T^*_xQ$. The action map $F\colon Q\times {\mathfrak g}\to TQ$ is defined as $F(x,\xi)=\xi_Q(x)$. Consider a Lagrangian on $Q$ with Legendre transformation ${\mathcal L} \colon TQ\to T^*Q$. Then $$ {\mathbb I}=({\mathcal L} F)^*J\colon Q\times {\mathfrak g}\stackrel{F}{\to} TQ\stackrel{{\mathcal L}}{\to} T^*Q \stackrel{J}{\to} {\mathfrak g}^* $$ is an exact abstract moment map. For example, assume that ${\mathcal L}$ arises from a Riemannian metric $\left<~,~\right>$ on $Q$ so that ${\mathcal L}(v)= \left<v,\cdot\right>$ for a tangent vector $v$. Then ${\mathbb I}^\zeta(x,\xi)=\left<\xi_Q(x),\zeta_Q(x)\right>$. The map ${\mathbb I}$, called the locked momentum map, is used in the analysis of relative equilibria. (See \cite{SLM} and \cite{Le}.) Note that in general $Q\times {\mathfrak g}$ is not a symplectic manifold and $G$, in this example, does not have to be commutative. \end{Example} \begin{Remark} In view of Example \ref{exam:presymplectic}, it is worth pointing out that a moment map on a Poisson manifold (see, e.g., \cite{wein-lect}) may not be an abstract moment map even when the Poisson structure is preserved by the action. The reason is that, since a moment map is defined only up to addition of Casimir functions, and since on a Poisson manifold Casimir functions often exist in abundance, a moment map on a Poisson manifold may not be constant on the fixed point set. \end{Remark} \section{Existence of abstract moment maps} \labell{sec:existence} Every manifold with a $G$-action admits an abstract moment map: the zero map. This map is never proper unless the manifold is compact. In this section we answer the question of when a $G$-manifold admits a proper (in fact, polarized, see Definition \ref{polarized}) abstract moment map. A necessary condition for an action to admit a proper abstract moment map $\Psi$ is that each component of the fixed point set be compact. (Recall that $\Psi$ is constant on each such component.) Is this condition sufficient? Moreover, does a (proper) abstract moment map exist with prescribed values at the fixed points? \subsection{Existence of abstract moment maps for circle actions} Answers to the above questions take a particularly simple and attractive form when $G$ is a circle, when abstract moment maps are simply $G$-invariant functions that are constant on the connected components of the fixed point set. \begin{Theorem} \labell{thm:circle-exist} Let $G$ be a circle acting on $M$, and let $\psi \colon M^G\to {\mathbb R}$ be a locally constant function. \begin{enumerate} \item There exists an abstract moment map $\Psi\colon M\to{\mathbb R}$ with $\Psi|_{M^G}=\psi$. \item Assume that $\psi$ is proper and bounded from below. Then $\Psi$ can be chosen to be proper and bounded from below. \end{enumerate} \end{Theorem} \begin{Remark} In other words, if $G$ is the circle group, we can prescribe the values of an abstract moment map on the connected components of $M^G$ completely arbitrarily. If $M^G$ is compact, the condition of the second assertion is satisfied automatically, and every locally constant function on $M^G$ extends to a proper abstract moment map. \end{Remark} \begin{proof} The theorem follows from the following two facts, applied to $X = M^G$ and $f = \psi$. \begin{enumerate} \item \labell{part 1} Let $X \subset M$ be a closed sub-manifold and $f \colon X \to {\mathbb R}$ a smooth function. Then there exists a smooth function $F \colon M \to {\mathbb R}$ such that $F|_X=f$. Moreover, if $f$ is bounded from below, $F$ can be chosen to be bounded from below too, and if $f$ is proper and bounded from below, $F$ can be chosen to be proper and bounded from below. \item \labell{part 2} Let $F \colon M \to {\mathbb R}$ be proper and bounded from below. Then the average $\overline{F}$ of $F$ by a compact group action is also proper and bounded from below. \end{enumerate} Let us prove the first fact. Fix a tubular neighborhood $U$ of $X$ in $M$, and let $\pi \colon U \to X$ be a smooth projection which extends to a proper map from the closure $\overline{U}$ to $X$. Let $\rho$, $1-\rho$ be a smooth partition of unity subordinate to the covering of $M$ by the two open sets $U$ and $M \smallsetminus X$. Pick a smooth function $\varphi \colon M \to {\mathbb R}$ which is proper and bounded from below (see, e..g., \cite{GP}, Chapter 1, Section 8). Then $F = \rho f + (1 - \rho) \varphi$ has the desired properties. To prove the second fact, notice that ${\overline{F}}^{-1}([-a,a])$ is contained in $G\cdot F^{-1}([-a,a])$, which is the image of the compact set $G\times F^{-1}([-a,a])$ under the continuous action mapping $G\times M\to M$. \end{proof} \begin{Remark} In Theorem \ref{thm:circle-exist} it is not true that if $\psi$ is just proper (but not bounded) then $\Psi$ can be chosen to be proper. In general, a proper map on a closed submanifold $X$ of $M$ might not extend to a proper map on $M$. For instance, the function $f(0,y) = y$ on the $y$-axis does not extend to a continuous proper function $F$ from ${\mathbb R}^2$ to ${\mathbb R}$. A similar counterexample involving abstract moment maps is given below. \end{Remark} \begin{Example} Let $M$ be obtained by the following plumbing construction: $$ M = {\mathbb Z} \times S^2 \times D^2 / \sim,$$ where $S^2 = \{ (x,y,z) \in {\mathbb R}^3 \ | \ x^2 + y^2 + z^2 =1 \}$, $D^2 = \{ (u,v) \in {\mathbb R}^2 \ | \ u^2 + v^2 < \epsilon^2 \}$, and $(n,x,y,\sqrt{1-x^2-y^2},u,v) \sim (n+1,u,v,-\sqrt{1-u^2-v^2},x,y)$ for all $n$. Take the diagonal circle action: $$ e^{i\theta} \cdot [n,x,y,z,u,v] = [n,x',y',z,u',v']$$ where $$ \left[ \begin{array}{cc} x' & u' \\ y' & v' \end{array} \right] = \left[ \begin{array}{cc} \cos\theta & \sin\theta \\ -\sin\theta & \cos\theta \end{array} \right] \left[ \begin{array}{cc} x & u \\ y & v \end{array} \right]. $$ The function \begin{equation} \labell{prop not pol} \psi([n,0,0,1,0,0]) = n \end{equation} is a locally constant function on the fixed point set and is proper, but it does not extend to a proper function $\Psi\colon M \to {\mathbb R}$. (The function $\psi$ extends to a (non-proper) Hamiltonian moment map, for a closed two-form whose pullback to each $\{ n \} \times S^2 \times \{ 0 \}$ is non-negative and has total area one.) \end{Example} \subsection{Assignments} Let us now investigate more closely the question of existence of an abstract moment map for an action of a torus whose dimension is greater than one. Theorem \ref{thm:circle-exist} is no longer true in this case: \begin{Example} \labell{exam:S2xS2} Let $\Psi = (\Psi_1,\Psi_2)$ be an abstract moment map on $M=S^2 \times S^2$ with $G = S^1 \times S^1$ acting by rotating each of the two factors. Then $\Psi$ must send the four fixed points to the corners of a rectangle in ${\mathbb R}^2$ whose sides are parallel to the axes. Thus the values $\Psi(M^G)$ cannot be assigned arbitrarily. \end{Example} Moreover, the abstract moment maps might not even separate the components of the fixed point set: \begin{Example} \labell{S4} Let $S^4$ be the unit sphere in ${\mathbb C}\times{\mathbb C}\times{\mathbb R}$, and let $G=S^1\times S^1$ act on it by rotating each of the the first two factors. There are two fixed points: the North Pole and the South Pole. The set of points fixed by the first $S^1$ is connected and contains both poles. The same is true for the set of points fixed by the second $S^1$. Consequently, any abstract moment map on $S^4$ must have the same value at the poles. \end{Example} These examples stress the role of the orbit type strata other than the components of the fixed point set. For a stratum $X$ we denote by ${\mathfrak g}_X$ the infinitesimal stabilizer of any of its points. Suppose that $\Psi \colon M \to {\mathfrak g}^*$ is an abstract moment map. Then for each infinitesimal orbit type stratum $X$ in $M$, the map $\Psi$ followed by the projection ${\mathfrak g}^* \to {\mathfrak g}_X^*$ gives an element $A(X)$ of ${\mathfrak g}_X^*$. An important observation is that the existence question for an abstract moment map is equivalent to the existence question for such an assignment, $X \mapsto A(X)$. We make this precise in Theorems \ref{thm:assignment} and \ref{thm:assignment proper}, which rely on the following definition and example. \begin{Definition} \labell{def-assign} An \emph{assignment} is a function $A$ that associates to each infinitesimal orbit type stratum $X$ in $M$ an element $A(X)$ of ${\mathfrak g}_X^*$ and that satisfies the following compatibility condition: if $X$ is contained in the closure of $Y$ then $A(Y)$ is the image of $A(X)$ under the restriction map ${\mathfrak g}_X^* \to {\mathfrak g}_Y^*$. The linear space of all assignments on $M$ is denoted by ${\mathcal A}(M)$. \end{Definition} In Section \ref{sec:cohomology} we discuss assignments in a broader, more abstract, context. \begin{Example} \labell{exam:moment-assign} Let $\Psi \colon M \to {\mathfrak g}^*$ be an abstract moment map. Then $A(X) = \Psi^{{\mathfrak g}_X}(X)$ is an assignment. The assignment $A$ and the moment map $\Psi$ are said to be associated with each other. If the abstract moment map is exact, i.e., $\Psi$ arises from a one-form $\mu$ so that $\Psi^\xi = \mu(\xi_M)$, the corresponding assignment is zero. \end{Example} \begin{Example} \labell{exam:assign-circle} When $G$ is the circle group, an assignment simply associates a real number to each component of the fixed point set. Thus, in this case, ${\mathcal A}(M)=({\mathfrak g}^*)^{\pi_0(M^G)}$. \end{Example} \begin{Example} \labell{CP2-T2} Consider the action of the two-dimensional torus $G=S^1 \times S^1$ on $M={\mathbb C}{\mathbb P}^2$ given by $(t_1,t_2)[z_0:z_1:z_2]=[z_0:t_1z_1:t_2z_2]$. This action has three fixed points, and every assignment $A$ is uniquely determined by its value at the fixed points. There are, however, three relations between the values $A|_{M^G}\in({\mathfrak g}^*)^3$, coming from the strata with one-dimensional stabilizers. As a result, ${\mathcal A}(M)$ is three-dimensional. Geometrically, the assignment values at the fixed points are the vertices of a triangle in ${\mathbb R}^2$ with two equal sides that are parallel to the coordinate axes. \end{Example} \begin{Example} \labell{exam:product} Let $M_1$ be a $G_1$-manifold and $M_2$ be a $G_2$-manifold. Then $$ {\mathcal A}(M_1\times M_2)={\mathcal A}(M_1)\oplus {\mathcal A}(M_2) $$ for the $G_1\times G_2$-action on $M_1\times M_2$. \end{Example} \begin{Remark} Replacing in Definition \ref{def-assign} the infinitesimal orbit type stratification by the orbit type stratification notion leads to the same class of assignments ${\mathcal A}(M)$. Namely, a function that associates to each orbit type stratum $X$ an element of ${\mathfrak g}_X^*$ and that satisfies the compatibility condition of Definition \ref{def-assign} is in fact constant on each infinitesimal orbit type stratum, and, hence, is an assignment. \end{Remark} \begin{Example} \labell{toric:1} Let $M$ be an $n$-dimensional complex manifold and let $G$ be an $n$-dimensional torus that acts on $M$. Suppose that each point of $M$ with stabilizer $H \subseteq G$ has a neighborhood which is biholomorphic to a neighborhood of the origin in ${\mathbb C}^n$ with an action of $H$ of the following form. The $H$-action is obtained as the composition of an isomorphism $H\to (S^1)^{\dim H}$ with the $(S^1)^{\dim H}$-action on ${\mathbb C}^{\dim H} \times {\mathbb C}^{n - \dim H}$ which is standard on the first factor and trivial on the second. (For instance, this is the case if $M$ is a toric manifold; see \cite{fulton} or \cite{audin}.) It follows that for any stratum $X$, the natural map $$ {\mathfrak g}_X^* \stackrel{\oplus_Y \pi^X_Y}{\longrightarrow} \bigoplus\limits_{\{ Y \ | \ X \preceq Y, \ \dim {\mathfrak g}_Y=1 \}} {\mathfrak g}_Y^* $$ is a linear isomorphism. Therefore, a moment assignment is determined by its values on the strata $Y$ with $\dim {\mathfrak g}_Y = 1$, and these values can be prescribed arbitrarily. So for such $M$, $$ {\mathcal A}(M) =\bigoplus_Y{\mathfrak g}_Y^*\cong {\mathbb R}^{\# \{Y \ | \ \dim {\mathfrak g}_Y =1 \} }. $$ \end{Example} \begin{Example} \labell{toric:2} Let $M$ be a K\"ahler toric manifold (also see Example \ref{toric:1}) with moment map $\Psi \colon M \to {\mathfrak g}^*$. The image $\Psi(M)$ is a convex polytope \cite{At:convexity,GS:convexity}. A convex polytope is stratified, with the strata being its open faces of various dimensions. The orbit type strata in $M$ are exactly the preimages in $M$ of the open faces in $\Phi(M)$, \cite{Del}. As a consequence, the poset of (infinitesimal) orbit type strata of $M$ is isomorphic to the poset of faces of $\Psi(M)$. Moreover, the stabilizers can be read from the faces and vice versa: the affine plane spanned by the face $\Psi(X)$ is a shift of the annihilator in ${\mathfrak g}^*$ of the Lie algebra ${\mathfrak g}_X$. The shifts are exactly given by the assignment $A(X) = \Psi^{{\mathfrak g}_X} (X)$: $$ \text{affine span} (\Psi(X)) = \text{ preimage of $A(X)$ under ${\mathfrak g}^* \to {\mathfrak g}_X^*$ }. $$ The polytope $\Psi(M)$, and hence the moment assignment $A$, determine the manifold, the $G$-action, and the symplectic form up to an equivariant symplectomorphism, \cite{Del}, and the equivariant K\"ahler structure on the strata, \cite{Gu}. \end{Example} \begin{Remark} In Example \ref{toric:2} we saw that the moment assignment of a symplectic toric manifold $M$ determines its moment polytope $\Psi(M)$. Similarly, for a toric manifold with a closed invariant two-form which may have degeneracies, the moment assignment determines its \emph{twisted polytope} in the sense of \cite{KT:toric}. \end{Remark} An obvious, but important, fact is \begin{Lemma} \labell{fix A} Let $\Psi_0$ and $\Psi_1$ be abstract moment maps which have the same assignment, $A$. Then $(1-\rho) \Psi_0 + \rho \Psi_1$ is also an abstract moment map with assignment $A$, for any invariant smooth function $\rho$. \end{Lemma} \begin{proof} For any $H \subseteq G$, on every component $X$ of $M^H$, we have $$ (1-\rho) \Psi_0^H + \rho \Psi_1^H = (1-\rho) A(X) + \rho A(X) \equiv A(X). $$ \end{proof} \begin{Remark} \labell{rmk:non-abelian} The definition of an assignment can be extended to non-commutative groups $G$. An assignment can then be defined as a function $x\mapsto A(x)\in {\mathfrak g}_x^*$ on $M$ such that the following conditions hold: \begin{itemize} \item $A(g\cdot x)=Ad^*_g A(x)$ for all $x\in M$ and $g\in G$. \item $A^{\mathfrak h}$ is locally constant on the set $M^{\mathfrak h}$ of points $x$ with ${\mathfrak h}\subseteq {\mathfrak g}_x$. \end{itemize} In the non-commutative case, as in Example \ref{exam:moment-assign}, an abstract moment map gives rise to an assignment. \end{Remark} \subsection{Existence of abstract moment maps for torus actions} \labell{subsec:polar} The relation between abstract moment maps and assignments is expressed in the following \begin{Theorem} \labell{thm:assignment} Let $M$ be a manifold with a $G$ action. Let $A: X \mapsto A(X)$ be an assignment. Then there exists an abstract moment map $\Psi \colon M \to {\mathfrak g}^*$ which is associated with $A$, i.e., such that $\Psi^{{\mathfrak g}_X}(X) = A(X)$ in ${\mathfrak g}_X^*$ for every orbit type stratum $X$. \end{Theorem} \begin{proof} Let $m$ be a point in $M$, let ${\mathfrak h}$ be the infinitesimal stabilizer of $m$, and let $A(m) \in {\mathfrak h}^*$ be the element assigned to the orbit type stratum containing $m$. Let $\Psi_m \in {\mathfrak g}^*$ be any element whose projection to ${\mathfrak h}^*$ is $A(m)$. Pick an open neighborhood $U_m$ of the orbit $G \cdot m$ which equivariantly retracts to the orbit. The constant function $\Psi_m$ is an abstract moment map on $U_m$ whose assignment is $A|_{U_m}$. Choose an invariant partition of unity $\{\rho_j\}$ subordinate to the covering of $M$ by the open subsets $U_m$, with the support of $\rho_j$ contained in the open set $U_{m_j}$. The convex combination $\Psi = \sum \rho_j \Psi_{m_j}$ is an abstract moment map; this follows from Lemma \ref{fix A}, applied to open subsets of the manifold. \end{proof} On a non-compact manifold, it is sometimes required that an abstract moment map be proper. (See \cite{Ka,GGK:lerman,GGK:book}.) In fact, we often need a component of $\Psi$ to be proper and bounded from below. \begin{Definition} \labell{polarized} Let $\eta \in {\mathfrak g}$ be a Lie algebra element. A function $\Psi \colon M \to {\mathfrak g}^*$ is said to be \emph{$\eta$-polarized} if its $\eta$th component, $\Psi^\eta \colon M \to {\mathbb R}$, is proper and bounded from below. \end{Definition} Note that an $\eta$-polarized function is necessarily proper, because its $\eta$-component is proper. However, not every proper map is $\eta$-polarized for some $\eta$. If $M$ is compact, $\Psi$ is automatically $\eta$-polarized for all $\eta$. Polarized abstract moment maps posses the following two properties, which may in generally fail for proper abstract moment maps: \begin{enumerate} \item A linear combination of $\eta$-polarized abstract moment maps on the same manifold is again an $\eta$-polarized (hence proper) abstract moment map. \item Let $\Psi_j\colon M_j \to {\mathfrak g}^*$, $j=1,2$, be $\eta$-polarized abstract moment maps. Consider the product $M_1 \times M_2$ with the diagonal $G$-action; let $\pi_1, \pi_2$ be the projection maps to $M_1$ and $M_2$. Then $\Psi_1 \circ \pi_1 + \Psi_2 \circ \pi_2$ is an $\eta$-polarized (hence proper) abstract moment map. \end{enumerate} Fix a $G$-manifold $M$ and a vector $\eta \in {\mathfrak g}$. Denote the set of zeros of $\eta_M$ by $M^\eta$. This is exactly the set of points whose infinitesimal stabilizer contains $\eta$. Therefore, the $\eta$-coordinate of any assignment is well defined on this set. \begin{Definition} An assignment $A$ is \emph{$\eta$-polarized} if its $\eta$-component on $M^\eta$, $$ A^\eta\colon M^\eta \to {\mathbb R},$$ is proper and bounded from below. \end{Definition} \comment{\ Omit the following remark. I'm keeping it in $\backslash$comment in case we want to use it for the book.\\ \begin{Remark} Let $H$ be the closure in $G$ of the one-parameter subgroup generated by $\eta$. Then the zero set $M^\eta$ of $\eta_M$ is equal to the fixed point set $M^H$ of $H$ in $M$. On this set, the $H$-component of an assignment is a well--defined function $A^H \colon M^H \to {\mathfrak h}^*$. An assignment $A$ is $\eta$-polarized if and only if this function is $\eta$-polarized in the sense of Definition \ref{polarized}. For a generic $\eta$, we have $H=G$. For such an $\eta$, an assignment is $\eta$-polarized if its restriction to the fixed point set, which is an ordinary function $A\colon M^G \to {\mathfrak g}^*$, is $\eta$-polarized in the sense of Definition \ref{polarized}. \end{Remark} } \begin{Theorem} \labell{thm:assignment proper} Let $M$ be a manifold with a $G$ action. For every $\eta$-polarized assignment $X \mapsto A(X)$ on $M$ there exists an $\eta$-polarized abstract moment map $\Psi \colon M \to {\mathfrak g}^*$ whose assignment is $A$. \end{Theorem} \begin{Corollary} \labell{cor:torus-exist} Assume that $M^G$ is compact. Then every assignment extends to a proper abstract moment map. \end{Corollary} As a consequence, if $M^G$ is compact, there always exists a proper abstract moment map (e.g., one which extends the zero assignment). \begin{proof}[Proof of Theorem \ref{thm:assignment proper}] The function $A^\eta\colon M^\eta \to {\mathbb R}$ is well defined, proper, and bounded from below. Since $M^\eta$ is closed, $A^\eta$ extends to a function $\varphi \colon M \to {\mathbb R}$ that is proper and bounded from below. (See item 1 in the proof of Theorem \ref{thm:circle-exist}.) For each $m \in M$, let $\Psi_m \in {\mathfrak g}^*$ be an element whose projection to ${\mathfrak g}_m^*$ is $A(m)$. We choose $\Psi_m \in {\mathfrak g}^*$ to meet the following additional requirement: $\Psi_m^\eta=\left< \Psi_m , \eta \right> = \varphi(m)$. If $m \in M^\eta$, this condition is automatically satisfied, and if $m \not \in M^\eta$, this choice is possible because $\eta\not\in{\mathfrak g}_m$. Let $U_m$ be a tubular neighborhood of the orbit through $m$ which equivariantly retracts to the orbit and on which the function $\varphi$ differs from the value $\varphi(m)$ by less than $1$. Then the constant function $\Psi_m$ is an abstract moment map on $U_m$ with assignment $A$ and whose $\eta$-component is bounded from below by $\varphi-1$. Choose an invariant partition of unity $\{\rho_j\}$ subordinate to the covering of $M$ by the open subsets $U_m$, with the support of $\rho_j$ contained in the open set $U_{m_j}$. Then the convex combination $\Psi = \sum \rho_j \Psi_{m_j}$ is an abstract moment map; this follows from Lemma \ref{fix A}, applied to open subsets of the manifold. Moreover, since the $\eta$-component of each $\Psi_m$ is bounded from below by $\varphi-1$, the same holds for $\Psi$. Since $\Psi^\eta \geq \varphi-1$, and $\varphi-1$ is proper and bounded from below, $\Psi^\eta$ is proper and bounded from below. \end{proof} In general, a $G$-manifold $M$ may admit no proper abstract moment maps, even when every connected component of the fixed point set $M^G$ is compact (and so a proper locally constant map $\psi\colon M^G\to{\mathfrak g}^*$ does exist). The obstruction lies in the compatibility condition; the manifold $M$ might not admit a proper assignment. We will now construct an example of such a $G$-manifold. \begin{Example} \labell{S4-chain} Let $G = S^1 \times S^1$ act on the four--dimensional sphere $S^4$ as in Example \ref{S4}. Recall that the fixed points are the North and South Poles and that any abstract moment map on $S^4$ must take the same value at these points. Fix some small $\epsilon>0$, and let $D^4$ be the $\epsilon$-ball in ${\mathbb C}\times{\mathbb C}$ with the $G$-action that rotates each of the two factors. Take the trivial disk bundle over $S^4$, \begin{eqnarray*} N & = & S^4\times D^4 \\ & = & \{ (z,z',x,w,w') \ | \ |z|^2 + |z'|^2 + x^2 =1 \text{ and } |w|^2 + |w'|^2 < \epsilon^2 \} \\ & \subset & {\mathbb C}^2 \times {\mathbb R} \times {\mathbb C}^2 \end{eqnarray*} with the diagonal action of $G$. Since the neighborhood of each of the two fixed points in $N$ is equivariantly diffeomorphic to $D^4\times D^4$, we can plumb an infinite sequence of such $N$'s. More explicitly, take $M= N\times{\mathbb Z}/\!\sim$ where the equivalence relation $\sim$ is \begin{equation} \labell{gluing-map} (z,z',x,w,w',n) \sim (w,w',-x,z,z',n+1) \end{equation} for all $x>0$ and $n \in {\mathbb N}$, whenever both $|z|^2+|z'|^2$ and $|w|^2+|w'|^2$ are less than $\epsilon$. Then $M$ is a $G$-manifold. The gluing map \eqref{gluing-map} reverses the orientation; however, we can get an orientation on $M$ by flipping the orientation of every other copy of $N$. An abstract moment map on $M$ must take a constant value on the infinite sequence of fixed points; such a map cannot be proper. \end{Example} \subsection{Minimal stratum assignments.} \labell{sec:minimal} Theorem \ref{thm:assignment} can be understood as that assignments are combinatorial counterparts of abstract moment maps. The amount of information needed to determine an assignment can be further reduced by taking a full advantage of the compatibility condition, as follows. Recall that the (infinitesimal) orbit type strata in $M$ are partially ordered; $X \preceq Y$ if and only if $X$ is contained in the closure of $Y$. The strata that are minimal under this ordering are exactly those that are closed subsets of $M$. The closure of any orbit type stratum in $M$ is a smooth sub-manifold which contains a minimal stratum. Every component of the fixed point set, $M^G$, is a minimal stratum. However, there can exist minimal strata outside the fixed point set $M^G$. Whether or not such strata exist is related to an algebraic property called \emph{formality}. Recall that a compact manifold $M$ is formal if one of the following equivalent conditions (see, e.g., \cite{Borel}, \cite{hsiang}, or \cite{Ki:book}) is satisfied: \begin{enumerate} \item $H^*_G(M)=H^*(M)\otimes H^*(BG)$ as an $H^*(BG)$-module; \item $H^*_G(M)$ has no $H^*(BG)$-torsion; \item the restriction $j^*\colon H^*_G(M)\to H^*_G(M^G)=H^*(M^G)\otimes H^*(BG)$ is a mono-morphism. \end{enumerate} Here is an interesting geometric consequence of formality: \begin{Proposition} \labell{exam:min-strata-formal} On a compact formal manifold $M$, every minimal stratum is a connected component of $M^G$. \end{Proposition} \begin{proof} Let $X$ be a minimal stratum and let $H$ be the connected component of identity of $G_x$ for $x\in X$. Assume $H \neq G$. Then the equivariant Thom class $\tau$ of the normal bundle to $X$ is a non-zero torsion element in $H^*(BG)$. In fact, $\tau$ is annihilated by the image of $H^*(B (G/H))\to H^*(BG)$. Alternatively, $j^*\tau=0$, because $X\cap M^G=\emptyset$. \end{proof} For example, when $M$ is compact symplectic with $G$ acting Hamiltonianly, $M$ is equivariantly perfect, and hence formal, (see \cite{Ki:book}), and the above analysis applies. In this case, however, to show that a minimal stratum $X$ consists of fixed points, it suffices to observe that $X$ is a compact symplectic manifold and $H$ acts Hamiltonianly on $X$, so $H$ must have fixed points on $X$. \comment{\ Remove the following Question. (Perhaps keep it as $\backslash$comment for our own use.)\\ \begin{Question} It would be interesting to find out to what extend the equivariant cohomology $H_G^*(M)$ can be recovered from the set of orbit type strata with its partial ordering and with their stabilizers or, alternatively, from the \emph{isotropy assignment} of Section \ref{subsec:other coefficients}.\\ The next question is a bit problematic: spectral sequences, etc. We need to think more of it before we ask it.\\ Moreover, can we reconstruct $H_G^*(M)$ if we know $H_G^*(X)$ for all the strata $X$, as well as the partial ordering between strata? \end{Question} Compare with what Kirwan does in her book. Also, compare with Bredon's theorem. He says something like that the equivariant homotopy type of a manifold is determined by the ordinary homotopy types of the strata and by their partial ordering.} \begin{Definition} A \emph{minimal stratum assignment} is an assignment of an element $A(X)\in{\mathfrak g}_X^*$ to each minimal stratum $X$, where ${\mathfrak g}_X$ is the infinitesimal stabilizer of $x\in X$, such that the following compatibility condition is satisfied: \emph{if two minimal strata $X_1$ and $X_2$ are such that $X_1\preceq Y$ and $X_2\preceq Y$ for some stratum $Y$, then the restrictions to ${\mathfrak g}_Y$ of $A(X_1)$ and of $A(X_2)$ are the same: $A(X_1)^{{\mathfrak g}_Y}=A(X_2)^{{\mathfrak g}_Y}$.} \end{Definition} Notice that this condition holds automatically for the zero assignment. The following theorem follows immediately from the definitions. \begin{Theorem} \labell{min-assign} The restriction of any assignment to the minimal strata is a minimal stratum assignment. Conversely, any minimal stratum assignment extends to a unique assignment. Hence, every minimal stratum assignment is associated with an abstract moment map. \end{Theorem} \begin{Remark} It appears that in Theorem \ref{min-assign} the minimal stratum assignment cannot be replaced by a function defined only on the fixed point set. Namely, we expect there to exist a $G$-manifold $M$ with isolated fixed points and a function $\psi \colon M^G \to {\mathfrak g}^*$ which does not extend to an assignment (hence does not extend to an abstract moment map), but which satisfies the following compatibility condition: if $x,y \in M^G$ belong to the same connected component of $M^H$, then $\psi^H(x) = \psi^H(y)$. \end{Remark} \begin{Question} In Remark \ref{rmk:non-abelian} we proposed a definition of assignments for an action of not necessarily abelian Lie group. It appears to be an interesting and feasible problem to check whether or not the results of this section generalize to such actions. \end{Question} \section{Exact moment maps} \labell{sec:exact} We have already shown that the natural forgetful homomorphism from the space of abstract moment maps on a $G$-manifold $M$ to the space ${\mathcal A}(M)$ of assignments on $M$ is onto (Theorem \ref{thm:assignment}). In this section we study the kernel of this epimorphism. Recall that an abstract moment map $\Psi$ is said to be exact if there exists a $G$-invariant one-form $\mu$ with $\Psi^\xi=\mu(\xi_M)$ for all $\xi\in{\mathfrak g}$ (see Example \ref{ex:one-form}). The assignment associated with such a map is zero. The following result, which is proved later in this section, shows that the converse is also true. \begin{Theorem} \labell{cor1} An abstract moment map whose assignment is identically zero is exact. More explicitly, suppose that $\Psi\colon M \to {\mathfrak g}^*$ is an abstract moment map such that for each subgroup $H \subset G$, the function $\Psi^H \colon M \to {\mathfrak h}^*$ vanishes on the $H$-fixed point set $M^H$. Then there exists an invariant one-form $\mu$ such that $\Psi^\xi = \mu(\xi_M)$ for all $\xi \in {\mathfrak g}$. \end{Theorem} Combining Theorems \ref{thm:assignment} and \ref{cor1} we obtain \begin{Corollary} \labell{cor12} The sequence $$ 0\to \left\{ \begin{array}{c} \text{exact}\\ \text{moment} \\ \text{maps} \end{array} \right\} \to \left\{ \begin{array}{c} \text{abstract} \\ \text{moment} \\ \text{maps} \end{array} \right\} \to {\mathcal A}(M)\to 0 $$ is exact. \end{Corollary} The proof of Theorem \ref{cor1} relies on the following key result, which we will prove in Section \ref{sec:proof}: \begin{Theorem} \labell{twoform-linear} Let $G$ be a torus acting linearly on ${\mathbb R}^m$, and let $\Psi$ be an abstract moment map on a neighborhood of the origin, vanishing at the origin. Then there exists a $G$-invariant one-form $\mu$ on a neighborhood of the origin such that $\mu(\xi_M)=\Psi^\xi$ for all $\xi \in {\mathfrak g}$. \end{Theorem} We will also need a parametric version of this theorem: \begin{Corollary} \labell{twoform-linear-par} Let $G$ be a torus acting linearly on the fibers of a vector bundle ${\mathcal V} \to Y$, and let $\Psi$ be an abstract moment map on a neighborhood of the zero section, vanishing on the zero section. Then there exists a smooth family $\mu$ of $G$-invariant one-forms on the fibers of ${\mathcal V}$, such that $\mu(\xi_M) = \Psi^\xi$ near the zero section. \end{Corollary} \begin{proof}[Proof of Corollary \ref{twoform-linear-par}] By using a partition of unity on $Y$, the corollary can be reduced to the case where $Y$ is a linear space and ${\mathcal V} = {\mathbb R}^m \times Y$. This case follows immediately from Theorem \ref{twoform-linear} when ${\mathbb R}^m$ is replaced by ${\mathbb R}^m\times Y$ with the trivial $G$-action on the second factor. \end{proof} Assuming Theorem \ref{twoform-linear}, let us prove a preliminary result, which is a local version of Theorem \ref{cor1} that will be used in the next section, and deduce Theorem \ref{cor1} from it. \begin{Proposition} \labell{prop:one-form} Let $G$ be a torus acting on a manifold $M$ and let $\Psi \colon M \to {\mathfrak g}^*$ be an abstract moment map. Let $p$ be a point in $M$ and $H = G_p$ its stabilizer. Suppose that $\Psi^H(p)=0$. Then there exists an open $G$-invariant neighborhood $V$ of $p$ in $M$ and a $G$-invariant one-form $\mu$ on $V$ such that \begin{equation} \labell{eq:one-form} \mu(\xi_M) = \Psi^\xi \end{equation} on $V$ for all $\xi \in {\mathfrak g}$. \end{Proposition} \begin{proof} Let us first examine the case where the action is locally free near $p$. Fix a basis $\xi_1,\ldots,\xi_n$ in ${\mathfrak g}$. Then the vector fields $(\xi_i)_M$ form a basis in the tangent space to the orbit at every point of a $G$-invariant neighborhood $V$ of the orbit through $p$. By setting \begin{equation} \labell{eq:loc-free} \mu((\xi_i)_M)=\Psi^{\xi_i} , \end{equation} we thus obtain a form defined along the orbits in $V$. We extend it to a differential form $\mu$ on $V$ by taking its composition with an orthogonal projection to the orbit with respect to a $G$-invariant metric. It is easy to see that $\mu$ satisfies the condition $\mu(\xi_M) =\Psi^\xi$. Let us now prove the proposition in the general case. Pick a closed subgroup $K \subset G$ whose Lie algebra ${\mathfrak k}$ is complementary to ${\mathfrak h}$ in ${\mathfrak g}$. A small $G$-invariant neighborhood $V$ of the orbit $Y$ through $p$ can be identified, by the slice theorem, with a neighborhood of the zero section in the normal bundle $ \pi \colon {\mathcal V} \to Y$ to $Y$ in $M$, with the action induced by that on $M$. We can apply Corollary \ref{twoform-linear-par} to the linear $H$-action on the fibers of ${\mathcal V}$, equipped with the abstract moment map $\Psi^H$ induced from $M$. As a result, we get a smooth family $\mu$ of one-forms on the fibers of ${\mathcal V}$, such that $\mu(\xi_M) = \Psi^\xi$ for all $\xi\in{\mathfrak h}$. The $K$-orbits form a foliation which is transverse to the fibration $\pi$. We extend $\mu$ to a one-form on a whole neighborhood of $Y$ by making $\mu$ vanish on the vectors tangent to the $K$ orbits. The resulting form is a $G$-invariant form $\mu_H$ on $V$ so that $\mu_H(\xi_M) = \Psi^\xi$ for all $\xi \in {\mathfrak h}$, and $\mu_H(\xi_M) = 0$ for all $\xi \in {\mathfrak k}$. The $K$-action on $V$ is locally free. Let $\mu_K$ be the form defined as above by \eqref{eq:loc-free} and extended to $V$ so that it vanishes on the vectors tangent to the fibers of $\pi$. Then $\mu_K(\xi_M) = \Psi^\xi$ for all $\xi \in {\mathfrak k}$. Since the vector fields $\xi_M$ for $\xi\in{\mathfrak h}$ are tangent to the fibers of $\pi$, we also have $\mu_K(\xi_M) = 0$ for all $\xi \in {\mathfrak h}$. The form $\mu=\mu_H+\mu_K$ has the desired property, that $\mu(\xi_M) = \Psi^\xi$ for all $\xi \in {\mathfrak g}$. \end{proof} \begin{proof}[Proof of Theorem \ref{cor1}] By Proposition \ref{prop:one-form}, there exists an open covering of $M$ by invariant sets $U_\alpha$, and on each $U_\alpha$ there exists an invariant one-form $\mu_\alpha$ such that $\Psi^\xi = \mu_\alpha(\xi_M)$ for all $\xi \in {\mathfrak g}$. Let $\rho_j$ be a partition of unity subordinate to this covering, with $\rho_j$ supported in $U_{\alpha_j}$ for each $j$. Define $\mu = \sum \rho_j \mu_{\alpha_j}$. Then $\Psi^\xi = \mu(\xi_M)$ on $M$ for all $\xi \in {\mathfrak g}$. \end{proof} \section{Hamiltonian moment maps} \labell{sec:forms} We have already seen (Example \ref{exam:presymplectic}) that every moment map which is associated with a closed invariant two-form is an abstract moment map. We will examine now the question of which abstract moment maps arise in this way. Recall that such abstract moment maps are called Hamiltonian. Thus we fix an abstract moment map $\Psi$, and we look for a closed two-form $\omega$ with $\iota(\xi_M) \omega = - d \Psi^\xi $. Note that such an $\omega$ would necessarily be $G$-invariant. Our first observation is an immediate consequence of the fact that every exact moment map, $\Psi^\xi=\mu(\xi_M)$, is automatically Hamiltonian with $\omega=d\mu$. Thus Theorem \ref{cor1} implies \begin{Corollary} \labell{cor11} Let $\Psi \colon M \to {\mathfrak g}^*$ be an abstract moment map with zero assignment. Then $\Psi$ is associated with an exact two-form. In particular, $\Psi$ is Hamiltonian. \end{Corollary} \subsection{Local existence of two-forms.} Our next result shows that there are no local obstructions to the existence of $\omega$, if $G$ is abelian. Quite surprisingly, a similar local existence result fails to hold for non-abelian compact groups, \cite{Br}. \begin{Corollary}[Local existence of two-forms] \labell{thm:presymplectic} Let $G$ be a torus acting on a manifold $M$, and let $\Psi \colon M \to {\mathfrak g}^*$ be an abstract moment map. For every $p\in M$, $\Psi$ is associated with an exact two-form $\omega$ on some open $G$-invariant neighborhood $V$ of $p$. In particular, $\Psi$ is Hamiltonian on a neighborhood of $p$. \end{Corollary} \begin{proof} Consider the new abstract moment map $\Psi - \Psi(p)$. By Proposition \ref{prop:one-form}, there exists an invariant neighborhood $V$ of $p$ in $M$ and a $G$-invariant one-form $\mu$ on $V$ such that $\mu(\xi_M) = \Psi^\xi - \Psi^\xi(p)$ on $V$. Let $\omega = d\mu$; then $\iota(\xi_M)\omega = -d\Psi^\xi$ on $V$. \end{proof} The following semi-local result is also of interest. \begin{Corollary} \labell{cor2} On a manifold with a unique minimal stratum, $X$, every abstract moment map is Hamiltonian. \end{Corollary} \begin{proof} Pick an element $\gamma \in {\mathfrak g}^*$ whose restriction to ${\mathfrak g}_X$ is equal to $\Psi^{{\mathfrak g}_X}(X)$, and apply Theorem \ref{cor1} to the abstract moment map $\Psi - \gamma$. \end{proof} \comment{\ Ax in the paper and keep for the book:\\ \begin{Corollary} In every $G$-manifold, there exists an invariant neighborhood of the fixed point set, $M^G$, on which every abstract moment map is Hamiltonian. \end{Corollary} } \begin{proof} Take a tubular neighborhood of $M^G$ which retracts to $M^G$, and apply Corollary \ref{cor2} to each of its connected components. \end{proof} \begin{Remark} It is well known that moment maps $\Psi$ associated with symplectic forms satisfy a certain non-degeneracy condition. For example, for circle actions the Hessian $d^2\Psi$ must be non-degenerate on the normal bundle to the fixed point set. In \cite{GGK:book}, we state explicitly a necessary and sufficient condition for $\Psi$ to be locally, near $M^G$, associated with a symplectic form. Furthermore, we will prove that abstract moment maps satisfying this non-degeneracy condition globally have many properties of moment maps on symplectic manifolds. These include the convexity theorem (\cite{At:convexity} and \cite{GS:convexity}) and formality (\cite{Ki:book}, see also Section \ref{sec:minimal} above). \end{Remark} \subsection{Global existence of two-forms} \labell{sec:forms-global} Let us now turn to the problem of global existence for $\omega$. The following example shows that not every abstract moment map is Hamiltonian. \begin{Example} \labell{CP2} Let $S^1$ act on ${\mathbb C}{\mathbb P}^2$ by $$ \lambda \cdot [z_0:z_1:z_2] = [z_0: \lambda z_1: \lambda^2 z_2]. $$ There are three fixed points: $[1:0:0]$, $[0:1:0]$, and $[0:0:1]$. Denote by $a,b,c$ their respective images by an abstract moment map. If the abstract moment map is associated with a closed two form, $\omega$, then it is an easy consequence of Stokes's theorem that the differences, $b-a$ and $c-b$ are, respectively, equal (up to a factor) to the integrals of $\omega$ on the 2-spheres $[*:*:0]$ and $[0:*:*]$ in ${\mathbb C}{\mathbb P}^2$. Since these lie in the same cohomology class, the values $a,b,c$ must then be equidistant: $a-b=b-c$. However, an abstract moment map can take arbitrary values $a,b,c$ at the three fixed points, by Theorem \ref{thm:circle-exist}. \end{Example} Recall that the equivariant cohomology classes in $H^2_G(M)$ are represented by the sums $\omega + \Psi$ where $\Psi$ is a Hamiltonian moment map and $\omega$ is a compatible two-form; see Example \ref{exam:presymplectic}. The forgetful mapping which sends $\omega+\Psi$ to the assignment $A$ corresponding to $\Psi$ gives rise to a homomorphism $$\rho\colon H^2_G(M)\to {\mathcal A}(M).$$ \begin{Theorem} \labell{thm:exist-2form} An abstract moment map $\Psi$ is Hamiltonian if and only if $A\in {\operatorname{im}}\, \rho$, where $A$ is the assignment of $\Psi$. \end{Theorem} \begin{proof} It is clear by definition that $A\in{\operatorname{im}}\,\rho$ if $\Psi$ is Hamiltonian. Conversely, assume that $A\in {\operatorname{im}}\,\rho$. Then there exists a $G$-equivariant equivariantly closed two-form $\omega+\Phi$ such that the assignment of $\Phi$ is also $A$. The difference $F=\Psi-\Phi$ is an abstract moment map with the zero assignment. By Theorem \ref{cor1}, $F$ is exact and therefore Hamiltonian (Corollary \ref{cor11}). Thus $\Psi$ is Hamiltonian as the sum of two Hamiltonian abstract moment maps, $F$ and $\Phi$. \end{proof} The space of Hamiltonian assignments, i.e., assignments associated with Hamiltonian abstract moment maps, is the quotient of the space of all Hamiltonian abstract moment maps by the space of exact abstract moment maps. This follows from Corollary \ref{cor12}. These three spaces fit together to form a part of a commutative square of exact sequences which summarizes some of our results. \begin{Proposition} \labell{prop:diagram} The following diagram is commutative and all of its rows and columns are exact: $$\begin{array}{ccccccccc} & & 0 & & 0 & & 0 & & \\ & & \downarrow & & \downarrow & & \downarrow & & \\ 0&\to & \left\{ \begin{array}{c} \text{basic}\\ \text{exact}\\ \text{2-forms} \end{array}\right\}& \to & \left\{\begin{array}{c} \mbox{equivariantly}\\ \mbox{exact}\\ \mbox{2-forms} \end{array}\right\}& \to & \left\{ \begin{array}{c} \mbox{exact} \\ \mbox{moment} \\ \mbox{maps} \end{array} \right\}& \to & 0\\ & & \downarrow & & \downarrow & & \downarrow & & \\ 0&\to & \left\{ \begin{array}{c} \text{basic}\\ \text{closed}\\ \text{2-forms}\end{array}\right\}&\to & \left\{\begin{array}{c} \mbox{equivariantly}\\ \mbox{closed}\\ \mbox{2-forms} \end{array}\right\}& \to & \left\{ \begin{array}{c} \mbox{Hamiltonian} \\ \mbox{moment} \\ \mbox{maps} \end{array} \right\}& \to & 0\\ & & \downarrow & & \downarrow & & \downarrow & & \\ 0 & \to & H^2(M/G) & \to & H^2_G(M) & \to & \left\{ \begin{array}{c} \mbox{Hamiltonian} \\ \mbox{assignments} \end{array}\right\}& \to & 0 \\ & & \downarrow & & \downarrow & & \downarrow & & \\ & & 0 & & 0 & & 0 & & \end{array} $$ \end{Proposition} \begin{proof} The exactness of the left column is a particular case of a more general fact, that the cohomology of the basic De Rham complex of $M$ is equal to $H^*(M/G)$, if $G$ is compact or, more generally, if the $G$-action is proper, even when the action is not free. This result, due to Koszul \cite{Koszul}, is similar to the De Rham theorem and can be proved in the same way. An easy proof is as follows. Recall that the sequence of sheaves of singular cochains on $M/G$ (with real coefficients) is a fine resolution of the constant sheaf ${\mathbb R}$ on $M/G$. Furthermore, basic forms on $G$-invariant open subsets of $M$ form a sheaf on $M/G$. This sheaf is a resolution of the locally constant sheaf because it is locally acyclic. Indeed, by using the fact that $G$ is compact (or that the action is proper) and adapting the proof of the Poincar\'{e} lemma, one can show that the basic cohomology of a neighborhood of an orbit is the same as of the orbit itself, i.e., zero in positive degrees. It is easy to see that this sheaf is also fine because it admits partitions of unity. Thus basic forms on $M$ provide another fine resolution of the constant sheaf on $M/G$. Since the cohomology of both resolutions are equal to the \v{C}ech cohomology of the constant sheaf on $M/G$, they are equal to each other. The middle column is exact by the definition of equivariant cohomology via the equivariant De Rham complex. (See, e.g., \cite{AB} and \cite{DKV}.) Exactness of the right column follows from Corollary \ref{cor12}. The fact that the top two rows are exact follows directly from the definitions of the spaces involved. The commutativity of the diagram is clear. Finally, commutativity with the exactness of the columns and the top two rows implies that the bottom row is exact by simple diagram chasing. \end{proof} \begin{Remark} Our notion of assignments has an interesting connection with a recent theorem of Goretsky-Kottwitz-MacPherson, \cite{GKM}. Assume that a compact oriented $G$-manifold $M$ is formal and satisfies in addition the so-called GKM condition, which we will recall below. Then the Goretsky-Kottwitz-MacPherson theorem implies that every assignment is Hamiltonian, and hence every abstract moment map is associated with a two-form. The ``GKM condition" is that the fixed points are isolated, and, additionally, every orbit type stratum with stabilizer of codimension one is two-dimensional. Let us now show how the above assertion follows from the theorem. Let $A$ be an assignment. Its restriction to the fixed point set is a locally constant function, $A^G\colon M^G \to {\mathfrak g}^*$. Such a function can be identified with an element of $H^2_G(M^G)$. For each subgroup $H \subset G$ of codimension one, the set of $H$-fixed points is a disjoint union of two-spheres in $M$, on each of which $G/H$ acts with exactly two fixed points; this is a consequence of the ``GKM condition" and formality. The assignment compatibility condition implies that in each such a two-sphere, the images in ${\mathfrak h}^*$ of $A^G$ are the same at the two fixed points. The Goretsky-Kottwitz-MacPherson theorem asserts that this condition on $A^G$ implies that there exists an equivariantly closed equivariant two-form, $\omega+\Psi$, on $M$, whose restriction to $M^G$ is $A^G$. By formality, an assignment on $M$ is uniquely determined by its restriction to $M^G$ (see Proposition \ref{exam:min-strata-formal}). Hence, $A$ is the assignment associated with $\Psi$; hence, it is Hamiltonian. \end{Remark} \comment{\ kill in the paper but keep for the book: As in Section \ref{sec:existence}, when $G$ is a circle, a necessary and sufficient condition for $\Psi$ to be associated with a closed two-form can be stated in terms of equivariant cohomology and the fixed point set. Recall that ${\mathcal A}(M)=({\mathfrak g}^*)^{\pi_0(M^G)}$ for $G=S^1$, i.e., ${\mathcal A}(M)$ is the set of functions from $\pi_0(M^G)$ to ${\mathfrak g}^*$ (Example \ref{exam:assign-circle}). Furthermore, since the $G$-action on $M^G$ is trivial, we have $H^2_G(M^G)=H^2(M^G)\oplus ({\mathfrak g}^*)^{\pi_0(M^G)}$, and thus $$ {\mathcal A}(M)=H^2_G(M^G)/H^2(M^G) .$$ Theorem \ref{thm:exist-2form} implies the following corollary which can also be proved directly: \begin{Corollary} \labell{thm:circle-exist-2form} Let $G$ be a circle. An abstract moment map $\Psi \colon M \to {\mathbb R}$ is Hamiltonian if and only if $\Psi|_{M^G}$ is in the image of the homomorphism $H^2_G(M)\to H^2_G(M^G)/H^2(M^G)$ \end{Corollary} \begin{Remark} When the fixed points of the action are isolated, $H^2(M^G)=0$ and the necessary and sufficient condition of Corollary \ref{thm:circle-exist-2form} simply turns into the condition that $\Psi|_{M^G}\in {\operatorname{im}}\, (H^2_G(M)\to H^2_G(M^G))$. Note also that even when $G$ is a torus, Corollary \ref{thm:circle-exist-2form} gives a necessary condition for $\Psi$ to be Hamiltonian. \end{Remark} The next example shows that Corollary \ref{thm:circle-exist-2form} does not extend to tori of dimension greater than one, i.e., the necessary condition is not then sufficient. \begin{Example} Let $\Psi \colon M \to {\mathfrak g}^*$ be an abstract moment map that is not Hamiltonian. (For instance, we can take $M = {\mathbb C}{\mathbb P}^2$ and $G=S^1$ as in Example \ref{CP2}.) Consider the product action of $G \times S^1$ on $M \times S^1$. Thus $G$ acts on the first factor $M$ and fixes the second factor $S^1$, and $S^1$ acts freely on the second factor $S^1$ and fixes the first factor $M$. Since there are no fixed points, the restriction of any abstract moment map to the fixed point set is trivially contained in the image of $H^2_{G \times S^1} (M \times S^1)$. We claim that $\tilde{\Psi} (p,a) := (\Psi(p),0)$ is an abstract moment map which is not Hamiltonian. To see that $\tilde{\Psi}$ is an abstract moment map, take any subgroup $H \hookrightarrow G \times S^1$. If the composition $H \to G \times S^1 \to S^1$ is not trivial, the fixed point set $M^H$ is empty, and we have nothing to check. Otherwise, $H$ is a subgroup of $G$. Thus $\tilde{\Psi}^H$ is equal to $\Psi^H$, and its restriction to $M^H$ is locally constant because $\Psi$ is an abstract moment map. If $\tilde{\Psi}$ were Hamiltonian, $\tilde{\Psi}^G = \Psi$ would also be Hamiltonian and so would be the restriction of $\Psi$ to $M \times \{ 1 \}$, contradicting the original assumption about $\Psi$. \end{Example} } \section{Abstract moment maps on linear spaces: \\ the proof of Theorem \ref{twoform-linear}.} \labell{sec:proof} A crucial step in the proof of Theorem \ref{cor1} and Proposition \ref{prop:one-form}, on which many of our subsequent results rely, is Theorem \ref{twoform-linear}. In this section we recall this theorem and prove it. A different proof can be found in \cite{GGK:book}. \smallskip \noindent \textbf{Theorem \ref{twoform-linear}.} \emph{ Let $G$ be a torus acting linearly on ${\mathbb R}^m$, and let $\Psi$ be an abstract moment map on a neighborhood of the origin, vanishing at the origin. Then there exists a $G$-invariant one-form $\mu$ on a neighborhood of the origin such that $\mu(\xi_M)=\Psi^\xi$ for all $\xi \in {\mathfrak g}$.} \begin{proof}[Proof of Theorem \ref{twoform-linear}] First note that by adding, if necessary, an additional copy of ${\mathbb R}$ (with the trivial $G$-action) to ${\mathbb R}^m$, we can always make $m$ even. There exists a $G$-invariant complex structure on the vector space ${\mathbb R}^m$; fix one. We obtain a representation $\rho$ of $G$ on ${\mathbb C}^d$ with weights $\alpha_1,\ldots,\alpha_d$. The infinitesimal action of the Lie algebra ${\mathfrak g}$ of $G$ on ${\mathbb C}^d$ is then given by the vector fields \begin{equation} \labell{eq:1} \xi_M = \sqrt{-1} \sum \limits_{i=1}^d \alpha_i(\xi) \left(z_i\frac{\partial}{\partial z_i}-\bar z_i\frac{\partial}{\partial\bar z_i}\right) . \end{equation} From now on we will forget about the reality conditions and assume that $\mu$ is a complex-valued one-form. When such a form is found, it will suffice to replace it by the real form $(\mu+\bar{\mu})/2$, which still has the desired properties because $\xi_M$ are real vector fields. We will also forget about the equivariance conditions. If a form $\mu$ such that $\mu(\xi_M) = \Psi^\xi$ for all $\xi \in {\mathfrak g}$ is constructed, and if $\Psi$ is equivariant, we can replace $\mu$ by its average. Any one-form on ${\mathbb C}^d$ can be written in the form \begin{equation} \labell{mu on Cd} \mu=-\sqrt{-1} \sum\limits_{i=1}^d f_i dz_i-g_i d\bar z_i \end{equation} for some smooth functions $f_j$ and $g_j$, $j=1,\ldots,d$. For such a one-form, the function $\Psi \colon {\mathbb C}^d \to {\mathfrak g}^*$ defined by $\Psi^\xi = \mu(\xi_M)$, with the vector fields $\xi_M$ given by \eqref{eq:1}, is \begin{equation} \labell{Psi on Cd} \Psi = \sum_{j=1}^d (z_j f_j + {\bar{z}}_j g_j) \alpha_j. \end{equation} Conversely, for any function $\Psi$ which has the form \eqref{Psi on Cd} for some smooth functions $f_j$, $g_j$, there exists a one-form $\mu$ such that $\Psi^\xi = \mu(\xi_M)$ for all $\xi \in {\mathfrak g}$. Namely, just take $\mu$ to be given by \eqref{mu on Cd}. To prove Theorem \ref{twoform-linear}, it is thus enough to prove the following \begin{Proposition} \labell{smooth} Let $\Psi$ be a ${\mathfrak g}^*$-valued function on a neighborhood of the origin in ${\mathbb C}^d$, vanishing at the origin and satisfying the second condition of an abstract moment map: \begin{quote} for any subgroup $H \subset G$, the function $\Psi^H \colon {\mathbb C}^d \to {\mathfrak h}^*$ is locally constant on the set of $H$-fixed points. \end{quote} Then there exist smooth functions $f_j$ and $g_j$ such that $\Psi$ is given by \eqref{Psi on Cd} on a neighborhood of the origin. \end{Proposition} \begin{Remark} The converse is easy: any $\Psi$ of the form \eqref{Psi on Cd} satisfies the second condition of an abstract moment map. \end{Remark} Let us start with polynomial functions and one-forms: \begin{Proposition} \labell{polynomial} Let $\Psi \colon {\mathbb C}^d \to {\mathfrak g}^*$ be a polynomial function which vanishes at the origin and which satisfies the second condition of an abstract moment map. Then there exist polynomials $f_j$ and $g_j$ on ${\mathbb C}^d$ such that $\Psi$ is given by \eqref{Psi on Cd}. \end{Proposition} \begin{proof}[Proof of Proposition \ref{polynomial}.] Since $\Psi$ is polynomial, we can write it uniquely as a sum of monomials, $$ \Psi = \sum\limits_{k,l} \beta_{k,l} z^k {\bar{z}}^l,$$ summing over $k = (k_1 , \ldots, k_d)$ and $l = (l_1,\ldots,l_d)$ in ${\mathbb N}^d$, where the coefficients $\beta_{k,l}$ are in ${\mathfrak g}^*$. For every subset $I \subset \{ 1, \ldots, d\}$, denote by $({{\mathbb C}^\times})^I$ the subset of ${\mathbb C}^d$ consisting of all vectors $(z_1,\ldots,z_d)$ for which $z_i \neq 0$ if and only if $i \in I$. All the points $z$ in $({{\mathbb C}^\times})^I$ have the same stabilizer, $G_I$, whose Lie algebra is \begin{equation} \labell{gI} {\mathfrak g}_I = \bigcap\limits_{i \in I} \ker \alpha_i. \end{equation} Since $\Psi$ satisfies the second condition of an abstract moment map, $\Psi^{{\mathfrak g}_I}$ is constant on $({{\mathbb C}^\times})^I$. Since, additionally, $\Psi$ is continuous and vanishes at the origin, $\Psi^{{\mathfrak g}_I}$ vanishes on $({{\mathbb C}^\times})^I$. Let us analyze what this condition tells us about the coefficients $\beta_{k,l}$. The polynomial $\Psi^\xi = \sum_{k,l} \beta_{k,l}(\xi) z^k {\bar{z}}^l$ vanishes on $({{\mathbb C}^\times})^I$ for all $\xi \in {\mathfrak g}_I$ if and only if for each $k,l$, the summand $\beta_{k,l}(\xi) z^k {\bar{z}}^l$ vanishes on $({{\mathbb C}^\times})^I$ for all $\xi \in {\mathfrak g}_I$. Fix $k$ and $l$, and restrict attention to $$I = I_{k,l} = \{ i \ | \ k_i \neq 0 \text{ or } l_i \neq 0 \}.$$ Since the monomial $z^k {\bar{z}}^l$ does not vanish on $({{\mathbb C}^\times})^I$, its coefficient, $\beta_{k,l} (\xi)$, must vanish for all $\xi \in {\mathfrak g}_I$. By \eqref{gI}, a linear functional that vanishes on ${\mathfrak g}_I$ is a linear combination of $\alpha_i$, $i \in I$. Therefore, $\beta_{k,l} = \sum\limits_{i \in I_{k,l}} \lambda_{i,k,l} \alpha_i$, and $$ \Psi = \sum_i \alpha_i \sum\limits_{k,l \text{ such that } i \in I_{k,l} } \lambda_{i,k,l} z^k {\bar{z}}^l.$$ Since for each $i \in I_{k,l}$, either $z_i$ or ${\bar{z}}_i$ factors out of the monomial $z^k {\bar{z}}^l$, $\Psi$ is of the form \eqref{Psi on Cd}. \end{proof} We will now show that the theorem we need to prove in the smooth category follows from its polynomial version, which has already been proved. In other words, we will deduce Proposition \ref{smooth} from Proposition \ref{polynomial}. To this end, let us reformulate these propositions as assertions that certain sequences of homomorphisms are exact. Denote by ${\mathcal P}$ the ring of complex-valued polynomials in $z_j$ and $\bar z_j$, $j=1,\ldots,d$. Define the modules ${\mathcal M}_i$, $i=1,\ 2, \ 3$, over ${\mathcal P}$ as follows: \begin{itemize} \item ${\mathcal M}_1$ is the space of one-forms $\sum f_i\,dz_i+g_i\,d{\bar z}_i$ with $f_i$ and $g_i$ in ${\mathcal P}$. \item ${\mathcal M}_2$ is the tensor product ${\mathcal P} \otimes {\mathfrak g}^*$ over ${\mathbb C}$. \item For each subset $I \subseteq \{ 1, \ldots, n \}$, denote by ${\mathcal P}_I$ the ring of polynomial functions on $({{\mathbb C}^\times})^I$. The restriction homomorphism ${\mathcal P} \to {\mathcal P}_I$ makes ${\mathcal P}_I$ into a ${\mathcal P}$-module. Set $$ {\mathcal M}_3 = \bigoplus_I {\mathcal P}_I \otimes {\mathfrak g}_I^* , $$ where ${\mathfrak g}_I$ is the Lie algebra of the stabilizer of $({{\mathbb C}^\times})^I$, given by \eqref{gI}. \end{itemize} Define the sequence of homomorphisms \begin{equation} \labell{eq:(2)} {\mathcal M}_1\overset{\alpha}{\to}{\mathcal M}_2\overset{\beta}{\to}{\mathcal M}_3 \end{equation} by setting $\alpha\colon \mu \mapsto \Psi$ with $\Psi^\xi=\mu(\xi_M)$ and $\beta$ to be the homomorphism $$ {\mathcal P} \otimes {\mathfrak g}^* \to \bigoplus_I {\mathcal P}_I \otimes {\mathfrak g}^*_I $$ associated with the restrictions ${\mathcal P} \to {\mathcal P}_I$ and ${\mathfrak g}^* \to {\mathfrak g}_I^*$. In other words, the $I$th component of $\beta(\Psi)$ is the ${\mathfrak g}_I^*$-component of $\Psi$ restricted to $({{\mathbb C}^\times})^I$. Hence, $\Psi$ satisfies the second condition of an abstract moment map if and only if $\beta(\Psi)=0$, and $\Psi$ is associated with a one-form if and only if it is in the image of $\alpha$. Proposition \ref{polynomial} is equivalent to the sequence \eqref{eq:(2)} being exact. Denote by ${\mathcal O}$ and ${\mathcal E}$, respectively, the algebras of germs of analytic, resp.\ smooth, functions on ${\mathbb C}^d$ at the origin. Let ${\mathcal M}^{{\mathit smooth}}_i$, where $i=1,\ 2,\ 3$, be the modules defined similar to ${\mathcal M}_i$ but in the category of smooth germs at the origin. Note that ${\mathcal M}^{{\mathit smooth}}_i$ are modules over ${\mathcal E}$. As before, we have a sequence of homomorphisms \begin{equation} \labell{exact-smooth} {\mathcal M}^{{\mathit smooth}}_1\overset{\alpha}{\to}{\mathcal M}^{{\mathit smooth}}_2\overset{\beta}{\to}{\mathcal M}^{{\mathit smooth}}_3. \end{equation} To prove the theorem in the smooth category, it suffices to show that this sequence is exact. Note that the inclusions ${\mathcal P}\to{\mathcal O}$ and ${\mathcal P}\to{\mathcal E}$ make ${\mathcal O}$ and ${\mathcal E}$ into ${\mathcal P}$-modules. The following lemma is obvious: \begin{Lemma} \labell{lemma:tensor} ${\mathcal M}^{{\mathit smooth}}_i={\mathcal M}_i\otimes_{{\mathcal P}}{\mathcal E}$. \end{Lemma} To finish the proof of Theorem \ref{twoform-linear}, we need to recall some facts from commutative algebra. Let $B$ be a commutative ring and $A$ a sub-ring of $B$. The ring $B$ is said to be {\em flat\/} over $A$ if for every exact sequence of $A$-modules $$ {\mathcal M}_1\to{\mathcal M}_2\to{\mathcal M}_3 $$ the sequence $$ {\mathcal M}_1\otimes_A B\to{\mathcal M}_2\otimes_A B\to{\mathcal M}_3\otimes_A B $$ is also exact. It is known that ${\mathcal O}$ is flat over ${\mathcal P}$ (see \cite{Mal}, page 45, Example 4.11) and ${\mathcal E}$ is flat over ${\mathcal O}$ (see \cite{Mal}, page 88, Corollary 1.2). This in turn implies that ${\mathcal E}$ is flat over ${\mathcal P}$. Therefore, the exactness of \eqref{eq:(2)} implies the exactness of \eqref{exact-smooth}. \end{proof} \section{Assignment cohomology} \labell{sec:cohomology} In this section we show that the space of assignments ${\mathcal A}(M)$ on a $G$-manifold fits as the zeroth space in a sequence of vector spaces $\operatorname{HA}^*(M)$, called the assignment cohomology. \subsection{Construction of assignment cohomology} Let $M$ be a manifold with an action of a torus $G$. Denote by $P_M$ the set of its infinitesimal orbit type strata (see Section \ref{sec:def-examples}). For each stratum $X$, denote by ${\mathfrak g}_X$ the infinitesimal stabilizer of the points of $X$, and let $$ V(X) = {\mathfrak g}_X^* $$ be the dual space. Recall that $P_M$ is a partially ordered set, a \emph{poset} for brevity, with $X \preceq Y$ if $X$ is contained in the closure of $Y$. Denote by $\pi_Y^X \colon V(X) \to V(Y)$ the natural projection dual to the inclusion map ${\mathfrak g}_Y \subseteq {\mathfrak g}_X$ when $ X \preceq Y$. Recall that the space of assignments is $$ {\mathcal A}(M)=\{ v \in \prod \limits_{X \in P_M} V(X) \ | \ \pi^X_Y v_X = v_Y \text{ for all } X \preceq Y \}. $$ Every abstract moment map induces an assignment. We will call elements of ${\mathcal A}(M)$ \emph{moment assignments}, to distinguish them from assignments with other coefficients $V(X)$, which we introduce later. Define the \emph{assignment cohomology} $\operatorname{HA}^*(M)$ to be the cohomology of the following cochain complex which we will denote by $C^*(M;V)$. A $k$-cochain is a function $\varphi$ that associates to each ordered $(k+1)$-tuple $X_0 \preceq \ldots \preceq X_k$ of elements of $P_M$ an element in $V(X_k)$. The differential $d$ is defined by the formula \begin{eqnarray} \labell{eq:diff} d\varphi(X_0,\ldots,X_{k+1}) &=&\sum _{\ell=0}^{k} (-1)^\ell \varphi(X_0,\ldots, \widehat{X_\ell},\ldots, X_{k+1}) \nonumber\\ &&\quad +(-1)^{k+1}\pi^{X_k}_{X_{k+1}}\varphi(X_0,\ldots, X_{k}) ,\end{eqnarray} where, as usual, the hat over $X_\ell$ means that $X_\ell$ is omitted. \begin{Example} The zeroth assignment cohomology is simply the space of assignments: $$ \operatorname{HA}^0(M) = {\mathcal A}(M).$$ \end{Example} \begin{Remark}[Functoriality] Assignment cohomology is functorial with respect to equivariant maps of manifolds: Let $M$ and $N$ be $G$-manifolds and let $f\colon M \to N$ be a $G$-equivariant map. Such a map might not send a stratum in $M$ to a stratum in $N$. (For example, the function $f(z,w) = z$ from ${\mathbb C}^2$ with the diagonal circle action to ${\mathbb C}$ with the standard circle action sends the open dense stratum ${\mathbb C}^2 \smallsetminus 0$ to the union of strata $\{ 0 \} \sqcup {\mathbb C}^\times = {\mathbb C}$.) However, it does induce a monotone mapping of posets, $\tilde{f} \colon P_M \to P_N$, in the following way. For each stratum $X$ in $M$ there exists a unique stratum $Y$ in $N$ with ${\mathfrak g}_{X} \subseteq {\mathfrak g}_{Y}$ whose closure contains $f(X)$. (To see this, consider the infinitesimal ${\mathfrak g}_X$-action in $N$. Since $X$ is connected, $f(X)$ is contained in a unique component of the ${\mathfrak g}_X$-fixed point set of $N$. This component is a $G$-invariant submanifold of $N$. The stratum $Y$ is the open dense stratum in this component.) We set $\tilde{f}(X) = Y$. Note that ${\mathfrak g}_X \subseteq {\mathfrak g}_{\tilde{f}(X)}$ for all $X$. By definition, the pullback map on the cochain complexes sends a cochain $\varphi \in C^k(N;V)$ to the cochain $f^*\varphi \in C^k(M;V)$ given by $$ (f^*\varphi)(X_0,\ldots,X_k) = \pi^{\tilde{f}(X_k)}_{X_k}\varphi(\tilde{f}(X_0),\ldots,\tilde{f}(X_k)). $$ The map $f^*$ commutes with $d$ and thus induces a pullback map in cohomology. \end{Remark} \begin{Theorem} \labell{vanishes} $\operatorname{HA}^k(M) = 0$ when $k \geq \dim M$ or $k \geq \dim G$. \end{Theorem} The proof of Theorem \ref{vanishes} will be an easy application of the following alternative construction of assignment cohomology. Let $C^k_0(M;V)$ be the space of functions $\varphi$ that associate to each ordered $(k+1)$-tuple $ X_0 \prec X_1 \prec \ldots \prec X_k$ of \emph{distinct} elements of $P_M$ an element of $V(X_k)$. This can be identified with the subspace of $C^k(M;V)$ consisting of those cochains that vanish on $(X_0,\ldots,X_k)$ whenever $X_i=X_{i+1}$ for some $i$. \begin{Theorem} \labell{C0} $C^*_0(M;V)$ is a subcomplex of $C^*(M;V)$, and the inclusion map of complexes induces an isomorphism in cohomology. \end{Theorem} This result is standard. However, for the sake of completeness, we prove it below. The complex $C^*_0(M;V)$ is much smaller than $C^*(M;V)$ and is more convenient to use for explicit calculations. Its disadvantage is that this complex is not functorial with respect to mappings of posets: a map $f \colon M \to N$ that send strata to strata sends a tuple $X_0 \prec \ldots \prec X_k$ to a tuple $f(X_0) \preceq \ldots \preceq f(X_k)$, but the $f(X_j)$ might not be distinct even if the $X_j$ are. \begin{proof}[Proof of Theorem \ref{vanishes}] By Theorem \ref{C0} it suffices to prove Theorem \ref{vanishes} for the cohomology of the complex $C^*_0(M;V)$. The first part of the theorem follows from the fact that if $X \prec Y$ then $\dim X < \dim Y$. Therefore, the longest possible tuple $X_0 \prec \ldots \prec X_k$ of distinct strata has $k = \dim M$. Thus $C^k_0(M;V) = 0$ for $k > \dim M$. When $k = \dim M$, the maximal stratum $X_k$ is the open dense stratum in $M$, on which $V(X_k) = 0$. As a result, $C^{\dim M}_0 (M;V) = 0$. The second part of the theorem follows from the fact that if $X \prec Y$ then $\dim {\mathfrak g}_X > \dim {\mathfrak g}_Y$. The same argument as before shows that $C^k_0(M;V) = 0$ whenever $k \geq \dim G$. \end{proof} Notice that the proof of the second part of the theorem breaks down if the infinitesimal orbit type stratification is replaced by the orbit type stratification. \begin{proof}[Proof of Theorem \ref{C0}] Denote by $(X_0^{k_0}, \ldots, X_l^{k_l})$ the tuple $$ (X_0,\ldots,X_0,\ldots,X_l,\ldots,X_l) $$ in which each $X_j$ occurs $k_j$ times and the strata $X_j$ are ordered and distinct: $X_0 \prec X_1 \prec \ldots \prec X_l$. Denote by $n(X_0^{k_0}, \ldots, X_l^{k_l})$ the number of $j$'s such that $k_j > 1$; call this number the \emph{fatness} of the tuple. As is easy to check, the fatness of a $(k+1)$-tuple is no greater than $k$. For each integer $n \geq 0$ let $C^k_n(M;V)$ be the space of $(k+1)$-cochains that are supported on tuples of fatness $n$. (This is consistent with the previous definition of $C^k_0(M;V)$.) Then $$ C^k(M;V) = C^k_0(M;V) \oplus \ldots \oplus C^k_k(M;V)$$ as vector spaces. Set $$ C^k_{>0}(M;V) = \bigoplus\limits_{n=1}^k C^k_n(M;V) .$$ For every nonzero cochain $\varphi$ in this space, there exists a unique integer $n$ between $1$ and $k$ and a unique decomposition \begin{equation} \labell{varphi1k} \varphi = \varphi_n + \varphi_{n+1} + \ldots +\varphi_k \end{equation} such that $\varphi_j \in C^*_j(M;V)$ for all $j$ and such that $\varphi_n \neq 0$. An easy computation shows that $d(C^k_n (M;V)) \subseteq C^{k+1}_n (M;V) + C^{k+1}_{n+1} (M;V)$ for all $n \geq 1$ and that $d(C^k_0 (M;V)) \subseteq C^{k+1}_0 (M;V)$. In particular, $C^*_0(M;V)$ and $C^*_{>0}(M;V)$ are subcomplexes, and the assignment cohomology splits: $$ \operatorname{HA}^*(M;V) = \operatorname{HA}^*_0(M;V) \oplus \operatorname{HA}^*_{>0}(M;V).$$ It remains to show that $\operatorname{HA}^*_{>0}(M;V)$ vanishes. Define a linear map $L \colon C^k(M;V) \to C^{k-1}(M;V)$ by $$ (L \varphi) (X_0^{k_0} , \ldots , X_l^{k_l}) = \sum_{j=0}^l (-1)^{k_0 + \ldots + k_{j-1}} \varphi(X_0^{k_0}, \ldots , X_j^{k_j+1}, \ldots ,X_l^{k_l}).$$ An explicit computation shows that \begin{equation} \labell{dL} dL\varphi + Ld\varphi = \pi \varphi, \end{equation} where $$(\pi \varphi) (X_0^{k_0} ,\ldots , X_l^{k_l}) = n(k_0,\ldots,k_l) \varphi(X_0^{k_0},\ldots,X_l^{k_l}).$$ Therefore, $\pi\colon C^*(M;V)\to C^*(M;V)$ is chain homotopic to zero and, as a consequence, induces the zero map $\pi_*$ in homology. Denote by $j\colon C^*_{>0}(M;V)\to C^*(M;V)$ the natural inclusion. It is clear that $\pi j\colon C^*_{>0}(M;V)\to C^*_{>0}(M;V)$ is an isomorphism. Thus $\pi_* j_*\colon \operatorname{HA}^*_{>0}(M;V)\to \operatorname{HA}^*_{>0}(M;V)$ is an isomorphism. Since $\pi_*=0$, this is possible only when $\operatorname{HA}^*_{>0}(M;V)=0$. \end{proof} \subsection{Assignments with other coefficients} \labell{subsec:other coefficients} The definitions of assignments and assignment cohomology extend word--for--word to other systems of coefficients. A system of coefficients $V$ on the poset $P_M$ is a function that associates a vector space $V(X)$ to each stratum $X$ and a linear map $ \pi^X_Y \colon V(X) \to V(Y)$ to each pair $X\preceq Y$, so that $\pi_Z^Y \pi_Y^X = \pi_Z^X$ whenever $X \preceq Y \preceq Z$ and $\pi_X^X = \operatorname{id}$. We define a differential complex $(C^*(M;V),d)$ as before and denote its cohomology by $\operatorname{HA}^*(M;V)$. A morphism $V_1 \to V_2$ of two systems of coefficients consists of a linear map $V_1(X) \to V_2(X)$ for each $X \in P_M$, such that the squares $$ \begin{array}{ccc} V_1(X) & \longrightarrow & V_2(X) \\ \downarrow & & \downarrow \\ V_1(Y) & \longrightarrow & V_2(Y) \end{array} $$ commute for all $X \preceq Y$. The systems of coefficients $V$ on $P_M$ form a category. The assignment cohomology groups, $\operatorname{HA}^n(M;V)$, and in particular the space of assignments ${\mathcal A}(M;V)$, are functorial in $V$. \begin{Remark} \labell{derived functors} One can think of the poset $P_M$ as a category in which there exists a single arrow $Y \to X$ whenever $X \preceq Y$. A system of coefficients is a contra-variant functor from $P_M$ to the category of vector spaces, and a morphism of systems of coefficients is a natural transformation. The space of assignments is the inverse limit: \begin{equation} \labell{liminv} {\mathcal A}(M;V) = {\lim\limits_{\longleftarrow}}_{X \in P_M} V(X), \end{equation} which is a functor in $V$.\footnote{\labell{foot:order} Note that the convention on the direction of morphisms commonly used to turn a poset into a category is opposite of the one employed in this paper. Alternatively, the poset of strata is sometimes given the order inverse of the one used above. However, the only essential point in the choice of directions of morphisms is that a poset should be made into a category so that $V$ becomes a \emph{contra-variant} functor.} A system of coefficients $V$ can be viewed as a \emph{pre-sheaf} on the category $P_M$, in the sense of \cite[Expose I, Definition 1.2]{SGA4}. (Also see \cite[Ch.~I \S 2 and \S 4]{Mo}.) The assignments are the global sections: \begin{equation} \labell{Gamma} {\mathcal A}(M;V) = \Gamma(P_M;V) , \end{equation} and the assignment cohomology is equal to the cohomology of this presheaf. More explicitly, the cohomology groups of the presheaf are defined to be the derived functors of the global section functor \eqref{Gamma} (as a functor in $V$); equivalently, the cohomology groups are the derived functors of the inverse limit functor \eqref{liminv}. The fact that these derived functors are the same as the cohomology of the complex $C^*(M;V)$ is Proposition 6.1 in \cite[ch.~II]{Mo}. \end{Remark} For some applications it is beneficial to work with systems of coefficients with values in categories other than the category of vector spaces. Let us illustrate this by some examples. \begin{Example} \labell{isotropy AA} Assignments with values in the functor \begin{equation} \labell{virtual reps} V(X) = \{ \text{ equivalence classes of representations of ${\mathfrak g}_X$ } \}. \end{equation} are called \emph{isotropy assignments}. A $G$-action gives rise to a canonical isotropy assignment defined as follows: to each $X$ associate the isotropy representation of ${\mathfrak g}_X$ on $T_pM$, $p \in X$. In a similar manner, any $G$-equivariant vector bundle over $M$ gives rise to an isotropy assignment. \end{Example} \begin{Example} For a symplectic manifold with a Hamiltonian torus action, its X-ray is, roughly speaking, the direct sum of the isotropy assignment and the moment map assignment. The notion of X-rays, introduced in \cite{tolman}, is central to the study of Hamiltonian torus actions. See \cite{tolman,hamiltonian,defone,metzler1,metzler2}. \end{Example} A broad class of systems of coefficients can be obtained by the following construction. Consider the category of sub-algebras of ${\mathfrak g}$ with morphisms given by inclusion maps. Let $V'$ be any functor from this category to the category of vector spaces (or modules, abelian groups, etc.). Then we get a system of coefficients with $V(X) = V'({\mathfrak g}_X)$. For instance, we get moment assignments from $V'({\mathfrak h}) = {\mathfrak h}^*$, and we get isotropy assignments from $V'({\mathfrak h}) = \{ \text{ virtual representations of } {\mathfrak h} \}$. \begin{Example} \labell{exam:min2} Assume that $M$ has a unique minimal stratum, $X_0$. This is the case, for instance, when $M$ is a vector space on which $G$ acts linearly. Then $\operatorname{HA}^0(M;V)=V(X_0)$ and $\operatorname{HA}^k(M,V)=0$ for all $k>0$. Indeed, the operator $$ (Q\varphi) (Y_1,\ldots,Y_k) = \varphi(X_0,Y_1,\ldots,Y_k) $$ satisfies $dQ\varphi+Qd\varphi=\varphi$, hence it is a homotopy operator for the complex $C^*(M;V)$. (Also see Remark \ref{rmk:min}.) \end{Example} Theorem \ref{C0} holds for assignment cohomology with many other systems of coefficients. For example, the theorem clearly holds whenever $V$ takes values in the category of vector spaces. When Theorem \ref{C0} applies, we also have the following variant of Theorem \ref{vanishes}: \begin{Proposition} \begin{enumerate} \item Let $V$ be such that $V(X)=0$ for the open stratum $X$. Then $\operatorname{HA}^k(M;V)=0$ when $k\geq \dim M$. \item Let $V$ be obtained as the pull-back of a functor on the sub-algebras of ${\mathfrak g}$ which vanishes on the zero sub-algebra. Then $\operatorname{HA}^k(M;V)=0$ when $k\geq \dim G$. \end{enumerate} \end{Proposition} The proof of this fact is entirely similar to the proof of Theorem \ref{vanishes}. \subsection{Assignment cohomology for pairs} Let $N$ be a subset of $M$ which is a union of strata. Define the \emph{relative} assignment cohomology $\operatorname{HA}^*(M,N;V)$ to be the cohomology of the sub-complex $C^*(M,N;V)$ of $C^*(M;V)$ formed by those cochains which vanish on all $(k+1)$-tuples $X_0 \preceq \ldots \preceq X_k$ in which all of the strata $X_j$ are in $N$. \begin{Theorem} \labell{thm:coh} There is a long exact sequence \begin{equation} \labell{eq:long-exact} \ldots\to \operatorname{HA}^*(M,N; V)\to \operatorname{HA}^*(M; V)\to \operatorname{HA}^*(N)\to \operatorname{HA}^{*+1}(M,N; V) \to \ldots , \end{equation} where the connecting homomorphism is given by the standard formula. \end{Theorem} \begin{proof} The theorem is an immediate consequence of the fact that the sequence of complexes \begin{equation} \labell{short exact} 0\to C^*(M,N;V)\to C^*(M;V)\to C^*(N;V)\to 0 \end{equation} is exact. To prove the exactness of \eqref{short exact}, note that the space $C^*(M,N;V)$ is the kernel of the restriction map $C^*(M;V) \to C^*(N;V)$ by its definition. To see that the restriction map is onto, note that every cochain $\varphi$ in $C^*(N;V)$ can be extended to a cochain $\tilde{\varphi}$ in $C^*(M;V)$ by declaring $\tilde{\varphi}(X_0,\ldots,X_k)$ to be zero whenever not all of $X_j$'s are in $N$. \end{proof} An alternative proof of Theorem \ref{thm:coh} in the case where $N$ is open is given in Remark \ref{rmk:pf-long-exact}. \begin{Remark} \labell{rmk:exact-seq'} It is not clear if there is a way to define relative assignment cohomology so that the sequence \eqref{eq:long-exact} is exact in the case where $N \subseteq M$ is $G$-invariant but not a union of strata. For instance, consider a $G$-manifold $B$ and let $G$ act on $M = B \times [0,1]$ by acting on the first factor. Let $N = B \times \{ 0, 1 \}$. Since every stratum in $M$ meets $N$, it seems reasonable to set $\operatorname{HA}^*(M,N)=0$. Then an exact sequence \eqref{eq:long-exact} would give an isomorphism between ${\mathcal A}(M)$ and ${\mathcal A}(N)$. However, these spaces are not isomorphic; in fact, ${\mathcal A}(N)$ is isomorphic to ${\mathcal A}(M) \oplus {\mathcal A}(M)$. \end{Remark} As is with other ``cohomology theories'', an exact sequence of coefficients gives rise to a long exact sequence in assignment cohomology: \begin{Theorem} \labell{short-long} A short exact sequence of systems of coefficients, \begin{equation} \labell{V123} 0 \to V_1 \to V_2 \to V_3 \to 0, \end{equation} induces a long exact sequence in cohomology, \begin{equation} \labell{HkV123} \to \operatorname{HA}^k(M,N;V_1) \to \operatorname{HA}^k(M,N;V_2) \to \operatorname{HA}^k(M,N;V_3) \stackrel{\delta}{\to} \operatorname{HA}^{k+1}(M,N;V_1) \to \end{equation} \end{Theorem} \begin{proof} The short exact sequence \eqref{V123} naturally induces a short exact sequence of complexes, $$ 0 \to C^*(M,N;V_1) \to C^*(M,N;V_2) \to C^*(M,N;V_3) \to 0,$$ and hence the long exact sequence \eqref{HkV123} in cohomology. \end{proof} \begin{Remark} \labell{M,N} Suppose that $N \subseteq M$ is an \emph{open} subset which is a union of strata. Note that being open is equivalent to the following condition: \begin{equation} \labell{N complete} \text{ for every pair } X \prec Y \text{ of strata in } M, \text{ if } X \subseteq N \text{ then also } Y \subseteq N. \end{equation} Then the relative assignment cohomology is equal to the ordinary assignment cohomology with a different system of coefficients: $$ \operatorname{HA}^n(M,N;V) = \operatorname{HA}^n(M;V_{M/N}) $$ where $V_{M/N}$ is the system of coefficients given by $$ V_{M/N}(X) = \begin{cases} V(X) & \text{if $X \not \subseteq N$},\\ 0 & \text{otherwise} \end{cases} $$ for any stratum $X$ in $M$, with the projection maps $$ (\pi_{M/N})^X_Y = \begin{cases} \pi^X_Y & \text{if both $X$ and $Y$ are not in $N$,}\\ 0 & \text{otherwise.} \end{cases} $$ The compatibility condition, \begin{equation} \label{compatibility} (\pi_{M/N})^Y_Z \circ (\pi_{M/N})^X_Y = (\pi_{M/N})^X_Z \text{ whenever } X \preceq Y \preceq Z, \end{equation} which is required for $V_{M/N}$ to form a system of coefficients, is satisfied if and only if $N$ meets the requirement \eqref{N complete}. \end{Remark} \begin{Remark} \labell{rmk:pf-long-exact} If $N \subseteq M$ is a union of strata and is open, the long exact sequence for the pair $(M,N)$ in Theorem \ref{thm:coh} follows from the long exact sequence for coefficients in Theorem \ref{short-long}. To see this, recall that $\operatorname{HA}^n(M,N;V) = \operatorname{HA}^n(M;V_{M/N})$ as explained in Remark \ref{M,N}. Furthermore, we set $V_N(X)=V(X)/V_{M/N}(X)$ with $(\pi_N)^X_Y=\pi^X_Y$ if $X$ and $Y$ are both in $N$ and $(\pi_N)^X_Y=0$ otherwise. Then we have $\operatorname{HA}^*(M;V_N)=\operatorname{HA}^*(N;i^*V)$, where $i \colon N \to M$ is the inclusion map. The sequence of systems of coefficients $ 0 \to V_{M/N} \to V \to V_N \to 0 $ is exact. By Theorem \ref{short-long}, this sequence gives rise to the long exact sequence which coincides with the sequence \eqref{eq:long-exact} of Theorem \ref{thm:coh}. \end{Remark} \begin{Remark} \labell{delta} Relative assignment cohomology is a sequence of functors $$ V \mapsto \operatorname{HA}^n(M,N;V) $$ from the category ${\mathcal C}_M$ of systems of coefficients on $P_M$ to the category of vector spaces. This sequence, together with the maps $\delta$ of \eqref{HkV123}, form a \emph{$\delta$-functor}. (Essentially, this means that short exact sequences in ${\mathcal C}_M$ induce long exact sequences in cohomology, as in Theorem \ref{short-long}. See \cite{lang}.) In the non-relative case $N=\emptyset$ this $\delta$-functor is \emph{universal}, i.e, the functors $V \mapsto \operatorname{HA}^n(M;V)$ are the derived functors of the assignment functor $V \mapsto {\mathcal A}(M;V)$; see Remark \ref{derived functors}. This remains true in the relative case if $N \subseteq M$ is a union of strata and is open: the relative assignment cohomology functors $V \mapsto \operatorname{HA}^n(M,N;V)$ are then the derived functors of the relative assignment functor $V \mapsto {\mathcal A}(M,N;V)$ which associates to each $V$ the space of assignments that vanish on $N$. However, for a general $N$, it is not clear if these functors are universal or, equivalently, whether or not $V\mapsto \operatorname{HA}^*(M,N;V)$ are the derived functors for $V\mapsto {\mathcal A}(M,N;V)$. \end{Remark} \begin{Remark} \labell{rmk:min} The poset of strata $P_M$ does not in general satisfy the following condition which is routinely required in some sources (e.g., \cite{Je}, \cite{Mas}, and \cite{Ru}): \begin{equation} \labell{eq:condP} \begin{array}{c} \text{for any } X \in P_M \text{ and } Y \in P_M \text{ there exists } Z \in P_M\\ \text{ such that } Z \preceq X \text{ and } Z \preceq Y. \end{array} \end{equation} (The reader should keep in mind that our order convention is opposite of the standard one; see footnote \ref{foot:order}.) This condition is met, for example, when $P_M$ has a minimal element, i.e., an $X_0 \in P_M$ such that $X_0 \preceq X$ for all $X \in P_M$. (Equivalently $X_0$ is a stratum which is contained in the closure of every stratum $X$.) With the condition \eqref{eq:condP}, the poset $P$ is called a \emph{directed set}. Under this condition, an inverse system $V$ of finite--dimensional vector spaces is automatically flabby (see, e.g., \cite{Je} and \cite{Ru}) and $\sideset{}{^{(k)}}\lim\limits_{\longleftarrow} V=0$ for all $k>0$. This generalizes Example \ref{exam:min2}. \end{Remark} \subsection{Examples of calculations of assignment cohomology} The following simple example shows that relative assignment cohomology can be non-trivial in degrees greater than zero. \begin{Example} \labell{exam:relative} Let $M={\mathbb C}{\mathbb P}^2$ and $G$ be the torus ${\mathbb T}^2$ acting on $M$ as in Example \ref{CP2-T2}, and $N=M^G$. Then ${\mathcal A}(M,N)=0$, ${\mathcal A}(N)=({\mathfrak g}^*)^3$ is six-dimensional, and $\dim{\mathcal A}(M)=3$. Furthermore, $\operatorname{HA}^{*>0}(M)=\operatorname{HA}^{*>0}(N)=0$. Thus \eqref{eq:long-exact} turns into the exact sequence $$ 0 \to {\mathcal A}(M) \to {\mathcal A}(N)\to \operatorname{HA}^1(M,N) \to 0 , $$ where $\dim \operatorname{HA}^1(M,N)=3$, as can also be checked by a direct calculation. \end{Example} \begin{Example}[Assignment cohomology for toric varieties] Let $M$ be a compact smooth K\"ahler toric manifold of complex dimension $n$ with moment map $\Psi \colon M \to {\mathfrak g}^*$. (See Examples \ref{toric:1} and \ref{toric:2}.) Recall that the poset $P_M$ of orbit type strata is isomorphic to the poset of faces $\Psi(X)$ of a simple polytope $\Psi(M)$, and for each stratum $X$, $$ \dim_{\mathbb C} X = \dim \Psi(X) = n - \dim {\mathfrak g}_X^*. $$ We will work with the system of coefficients $V(X) = {\mathfrak g}_X^*$. The zeroth assignment cohomology is the space of assignment which was computed in Example \ref{toric:1}. Namely, $$ \operatorname{HA}^0(M;V)=\bigoplus_Y{\mathfrak g}_Y^*, $$ where the summation is over all strata $Y$ with $\dim {\mathfrak g}_Y=1$. These strata correspond to the $(n-1)$-dimensional faces of $\Psi(M)$. In particular, $$ \dim \operatorname{HA}^0(M;V) = \text{the number of facets of $\Psi(X)$}. $$ Let us prove that the higher cohomology groups vanish: \begin{equation} \labell{higher vanish} \operatorname{HA}^k(M;V) = 0 \quad \text{for all $k \geq 1$.} \end{equation} By Theorem \ref{C0} it is enough to work with the complex $C^*_0(M;V)$. For a closed cochain $\varphi \in C^k_0(M;V)$, $k \geq 1$, we will find a primitive $(k-1)$-cochain $\psi \in C^{k-1}_0(M;V)$, i.e., a cochain $\psi$ such that $d\psi=\varphi$. Let $X_0 \prec \ldots \prec X_{k-1}$ be any ordered $k$-tuple of distinct strata. Recall that the natural map \begin{equation} \labell{projections} V(X_{k-1}) \stackrel{\oplus \pi^{X_{k-1}}_{X_k}}{\longrightarrow} \bigoplus\limits_{X_k}V(X_k), \end{equation} where $X_k$ is such that ${\operatorname{codim}} X_k=1$ and $X_{k-1}\preceq X_k$, is a linear isomorphism. Therefore, to define the value $\psi(X_0,\ldots,X_{k-1})$, which is an element of $V(X_{k-1})$, it is enough to specify the projections of these elements to all of the spaces $V(X_k)$ with $X_k$ as above. We require these projections to be \begin{equation} \labell{pi psi} \pi^{X_{k-1}}_{X_k} \psi(X_0,\ldots,X_{k-1}) = (-1)^k\varphi(X_0, \ldots, X_k). \end{equation} Let us show that $d\psi=\varphi$. Set $\varphi' = \varphi - d\psi$. Note that $\psi(X_0,\ldots,X_{k-1})=0$ when ${\operatorname{codim}} X_{k-1}=1$. Then it follows from the definition \eqref{pi psi} of $\psi$ and the definition \eqref{eq:diff} of the differential that $\varphi'$ vanishes on all tuples $X_0 \prec \ldots \prec X_k$ in which ${\operatorname{codim}} X_k = 1$. Again, by \eqref{eq:diff}, $d\varphi'(X_0,\ldots,X_{k+1}) = \pi^{X_k}_{X_{k+1}} \varphi'(X_0,\ldots,X_k)$ for all tuples $X_0 \prec \ldots \prec X_k \prec X_{k+1}$ in which ${\operatorname{codim}} X_{k+1} = 1$. Since $d\varphi'=0$, and since \eqref{projections} is a linear isomorphism, this implies that $\varphi' \equiv 0$. Hence, $\varphi = d\psi$ is exact. \end{Example} We now give an example of a manifold which has a non-trivial (absolute, not relative) first assignment cohomology. \begin{Example} Let $M = S^2 \times S^2 \times S^2$, and let $G = S^1 \times S^1$ act by $$ (a,b) \cdot (u,v,w) = (a \cdot u, b \cdot v, ab^{-1} \cdot w) $$ where on the right the dot denotes the standard $S^1$ action on $S^2$ by rotations. The moment assignments can be drawn as pictures showing the moment map images of the orbit type strata (the ``x-ray"). Such a picture is shown in Figure \ref{fig:example}. Notice that this picture is two--, not three--, dimensional. This arrangement can be moved around as long as the edges are shifted but not rotated. An assignment is therefore determined by the location of the bottom left vertex and the lengths of the three edges coming out of it. Therefore, \begin{equation} \labell{five} \dim \operatorname{HA}^0(M;V) = \dim {\mathcal A}(M;V) = 5. \end{equation} We will find the dimension of the first assignment cohomology space by using the Euler characteristic of the complex $C^*_0(M;V)$ of Theorem \ref{C0}. We have \begin{equation} \labell{euler} \begin{array}{ccl} \dim \operatorname{HA}^0(M;V) - \dim \operatorname{HA}^1(M;V) & = & \sum_k (-1)^k \dim C^k_0(M;V) \\ & = & \dim C^0_0(M;V) - \dim C^1_0(M;V) \end{array} \end{equation} because $C^k_0(M;V) = 0$ for all $k \geq 2$ (see also Theorem \ref{vanishes}). A $0$-cochain associates to each vertex an element of a two--dimensional space and to each edge an element of a one--dimensional space. Therefore, \begin{equation} \labell{28} \begin{array}{ccl} \dim C^0_0 (M;V) & = & 2 (\text{number of vertices}) + (\text{number of edges}) \\ & = & 2 \cdot 8 + 12 \\ & = & 28. \end{array} \end{equation} A $1$-cochain associates an element of a one dimensional space to each pair consisting of a vertex and an edge coming out of it. Therefore \begin{equation} \labell{24} \dim C^1_0 (M;V) = 24. \end{equation} Substituting \eqref{28}, \eqref{24}, and \eqref{five} in \eqref{euler}, we get $$ 5 - \dim \operatorname{HA}^1(M;V) = 28 - 24, $$ hence $$ \dim \operatorname{HA}^1(M;V) = 1. $$ \end{Example} \begin{figure} \begin{center} \setlength{\unitlength}{0.00083333in} \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% {\renewcommand{\dashlinestretch}{30} \begin{picture}(2124,1539)(0,-10) \path (12,12) (912,12) (912,912) (12,912) (12,12) \path (1212,612) (2112,612) (2112,1512) (1212,1512) (1212,612) \path (12,12) (1212,612) \path (912,12) (2112,612) \path (912,912) (2112,1512) \path (12,912)(1212,1512) \end{picture} } \end{center} \caption{A manifold with nonzero first assignment cohomology} \labell{fig:example} \end{figure} \subsection*{Generalizations of assignment cohomology} The results and definitions of this section can be generalized or altered in many natural ways. For instance, the assignment cohomology can be defined for an arbitrary system of coefficients with values in an abelian category on an arbitrary poset. In particular, in a more geometrical realm, the infinitesimal orbit type stratification can be replaced by the orbit type stratification and $V$ can then be the pull-back of a contra-variant functor on subgroups of $G$. Furthermore, instead of working with functors with values in finite--dimensional vector spaces one may consider functors with values in abelian groups or graded vector spaces or rings. In fact, such functors do arise arise in the study of symplectic manifolds, and ${\mathcal A}(M;V)$ can be viewed as a repository containing many of the invariants of the action. One important example is the isotropy assignments of Example \ref{isotropy AA} above. Moreover, most of the results of this section extend with obvious modifications to actions of finite or compact \emph{non-abelian} groups. For example, we can take $V$ to be the pull-back of a functor on the sub-algebras of ${\mathfrak g}$ which is invariant under conjugations. However, the space of moment map assignments for actions of non-abelian groups (Remark \ref{rmk:non-abelian}) does not arise as the zeroth cohomology groups of this type. In the non-abelian case, the space of moment map assignments does not seem to be associated with a functor on the poset of strata. Instead, to obtain this space one should work with the singular foliation of $M$ given by the decomposition of $M$ according to actual stabilizers, but not just their conjugacy classes. This renders a correct generalization of assignment cohomology to actions of non-abelian groups much less straightforward. \begin{Remark} The assignment cohomology appears to be related to Bredon's equivariant cohomology, \cite{Bredon}, and, perhaps, to Borel's equivariant cohomology with twisted coefficients. The nature and explicit form of these relations are, however, unclear to the authors. \end{Remark}
\section{Introduction} Neutrino masses and oscillations\cite{Pontec,MNS} belong to the most exciting area of particle physics. It is exactly this area where the Standard Model of quarks and leptons is challenged in a most successful way. The recent discovery of muon neutrino oscillations\cite{SuperK98a} and the evidence for solar neutrino deficits\cite{solarEx} in comparison to the Standard Solar Model\cite{Bahcall} demonstrate that the neutrinos are massive. Phenomenological analyses favor two solutions of the solar and atmospheric neutrino problems, see e.g.\cite{FLMS}--\cite{BW99a} and for a recent review see \cite{BiGG}: \begin{itemize} \item large $\nu_\mu -\nu_\tau$ mixing for atmospheric anomalies and matter enhanced\cite{MSW} small mixing angle oscillations (SMA MSW) for solar neutrinos, see e.g.\cite{SMAMSW}. \item vacuum oscillations and bimaximal\cite{BPWW98}--\cite{BHKR} or nearly bimaximal\cite{nearly-bimax} mixing of three light neutrinos. \end{itemize} We believe that one of these two possibilities is true although other options are not completely excluded, see e.g.\cite{BKS,Langacker,BiGG,Str99}. In a near future experimental information from SuperKamiokande and SNO will select the best scenario. In fact a recent measurement of day--night asymmetries\cite{day-night} imposes constraints on SMA MSW solutions. On the other hand the vacuum oscillation solutions may be confirmed or excluded by final results on seasonal effects and recoil energy distributions, see \cite{BFLi,BW99a,GKK} and references therein. A preliminary data from SuperKamiokande\cite{prelSK} are analyzed in Ref.\cite{BW99a}. In this article we discuss some consequences of the bimaximal mixing scenario. Evidently solutions of `just so' type\cite{just-so} are preferred by the existing data. However we do not exclude a possibility\cite{BGG} that a larger range of $\Delta m^2_\odot$ is allowed for solar neutrinos if errors are underestimated in one of the radiochemical measurements. In such a case `just so' fits are also better\cite{BFLi,BW99a}. In the following discussion we use as input parameters\cite{Suzuki99} $$\Delta m^2_\odot = 4.3\times 10^{-10}\ \mbox{eV}^2$$ for oscillations of solar and $$ 1.5\times 10^{-3}\ \mbox{eV}^2 \le \Delta m^2_{atm} \le 6\times 10^{-3}\ \mbox{eV}^2$$ for oscillations of atmospheric neutrinos. \section{Maki--Nakagawa--Sakata lepton mixing matrix} The Dirac masses of quarks and charged leptons exhibit a strongly ordered hierarchical structure $$ m_u \ll m_c \ll m_t\ , \quad m_d \ll m_s \ll m_b\ , \quad m_e \ll m_\mu \ll m_\tau \ .$$ We assume a similar hierarchical structure for the Dirac masses of neutrinos. However this hierarchical structure is drastically modified by huge Majorana masses of righthanded neutrinos. As a result of the seesaw mechanism\cite{seesaw} three light and nearly lefthanded neutrinos appear in the mass spectrum. Their Majorana masses can form patterns very different from the hierarchical one. For the same reason the structures of quark and lepton mixing matrices are also quite different. For quarks the Cabibbo-Kobayashi-Maskawa matrix\cite{CKM} $V_{CKM}$ is nearly diagonal and its largest off-diagonal elements \begin{equation} | V_{us}| \simeq | V_{cd} | \approx \theta_c \approx 0.22 \end{equation} are fairly small. All other off-diagonal elements of $V_{CKM}$ are much smaller and can be parameterized by higher powers of the Cabibbo angle $\theta_c$~\cite{WolfCKM}. $V_{CKM}$ can be written as a product of two matrices describing unitary transformations of quarks with weak isospin projections $I_3 = + 1/2$ ($u,c,t$) and $I_3 = - 1/2$ ($d,s,b$) respectively: \begin{equation} V_{CKM} = {V_{+}}^\dagger\, V_{-} \ \ . \end{equation} In the Standard Model only $V_{CKM}$ is observable. In models of quark masses and mixing both matrices $V_{\pm}$ are specified and related to Yukawa couplings of up and down type quarks. A particularly interesting approach\cite{Fritzsch} relates angles in $V_{+}$ and $V_{-}$ to mass ratios of up and down type quarks respectively. In this way a relation is derived for the mixing between the first two generations \begin{equation} \theta_c = \left| \sqrt{{m_d/ m_s}} + e^{i\phi} \sqrt{{m_u/ m_c}} \right| \approx \sqrt{{m_d/ m_s}}\ \ . \end{equation} The mixing angle $\theta_c$ is dominated by the contribution from the $I_3 = - 1/2$ sector. The flavor mixing matrix\cite{MNS} for three light neutrinos relates the neutrino mass and flavor eigenstates: \begin{equation} \left( \matrix{ \nu_e \cr \nu_\mu \cr \nu_\tau \cr}\right) = \left( \matrix{ U_{e1} & U_{e2} & U_{e3} \cr U_{\mu 1} & U_{\mu 2} & U_{\mu 3} \cr U_{\tau 1}& U_{\tau 2} & U_{\tau 3} \cr } \right) \left(\matrix{ \nu_1 \cr \nu_2 \cr \nu_3 \cr}\right) = U_{MNS} \left(\matrix{ \nu_1 \cr \nu_2 \cr \nu_3 \cr}\right) \ \ . \label{eq:MNS} \end{equation} $U_{MNS}$ can be written as a product of two matrices $U_{\pm}$ describing transformations of the neutrinos ($I_3 = +1/2$) and the charged leptons ($I_3 = -1/2$): \begin{equation} U_{MNS} = {U_{-}}^\dagger\, U_{+}\quad . \label{eq:MNS1} \end{equation} It is plausible that the pattern of mixing angles in $U_{-}$ describing the charged lepton sector resembles that in $V_{-}$ for the down type quarks, i.e. there is an appreciable mixing in the electron--muon plane whereas the other mixing angles are much smaller \begin{equation} U_{-} \approx \left(\matrix{ c & -s & 0 \cr s & c & 0 \cr 0 & 0 & 1 \cr}\right) \label{eq:U-1/2} \end{equation} where\cite{Fritzsch} \begin{equation} |s| \simeq \sqrt{m_e/m_\mu} \simeq \theta_c/3 \simeq 0.07 \qquad {\rm and} \qquad c = \sqrt{ 1 - s^2} \approx 1 - s^2/2 \ \ . \label{eq:s} \end{equation} The second equality in (\ref{eq:s}) follows from grand unification relations\cite{GJ79} \begin{equation} m_d \simeq 3m_e\ \ , \qquad m_s \simeq m_\mu/3 \ \ . \end{equation} In the sector $I_3 = + 1/2$ the situation is quite different. The structure of the matrix $U_{+}$ is not directly related to the Dirac mass matrix $\bf m_{_D}$ for neutrinos. It is strongly affected by the structure of the Majorana mass matrix $\bf M_{_R}$ for the righthanded neutrinos. As a consequence the structure of $U_{+}$ is different from the structure of $V_{+}$. We assume that the mixing of solar and atmospheric neutrinos is close to bimaximal, \begin{equation} U_{+} \approx U_{bm} = \left(\matrix{ 1/\sqrt{2} & 1/\sqrt{2} & 0 \cr - 1/2 & 1/2 & 1/\sqrt{2} \cr 1/2 & - 1/2 & 1/\sqrt{2} \cr}\right)\ \ . \label{eq:U+1/2} \end{equation} We also assume that the dominant contribution to deviations from the ideal bimaximal mixing originates from the sector $I_3 = - 1/2$, i.e. from the form of the matrix $U_{-}$. (Later in this paper we will demonstrate simple seesaw models in which corrections to eq.(\ref{eq:U+1/2}) are fairly small.) The main goal of the present paper is to describe phenomenological consequences of these two assumptions. For the sake of simplicity we neglect CP violation and consider all elements of $U_{MNS}$ as real. Then the lepton mixing matrix is completely specified. From eqs.(\ref{eq:MNS1}), (\ref{eq:U-1/2}) and (\ref{eq:U+1/2}) we obtain \begin{equation} U_{MNS} \approx \left( \matrix{ c/\sqrt{2} -s/2\ , & c/\sqrt{2} + s/2\ , & s/\sqrt{2} \cr -c/2 - s/\sqrt{2}\ ,& c/2 - s/\sqrt{2}\ , & c/\sqrt{2} \cr 1/2 & - 1/2 & 1/\sqrt{2} \cr} \right) \label{eq:MNS2} \end{equation} with $s$ and $c$ given in eq.(\ref{eq:s}). \section{Neutrino oscillations} It is well known\cite{BiG98,BW98a} that for $\Delta m^2_\odot \ll \Delta m^2_{atm}$ and small $|U_{e3}|$, oscillations of solar neutrinos effectively decouple from oscillations in long baseline (LBL) experiments and those of atmospheric neutrinos. Both types of oscillations can be described by effective two--neutrino oscillations characterized by phase differences \begin{equation} \Delta_\odot = { \Delta m^2_\odot L\over 4E} \label{eq:Delsun} \end{equation} for solar, and \begin{equation} \Delta_{atm} = { \Delta m^2_{atm} L\over 4E} \label{eq:Delatm} \end{equation} for LBL and atmospheric oscillations. In the above equations $L$ is the distance traveled by a neutrino of energy $E$. Probabilities of transitions between different neutrino flavors in atmospheric and LBL experiments are given by, see \cite{BiGG} and references therein: \begin{eqnarray} {\cal P}_{LBL}\left(\nu_\alpha\rightleftharpoons\nu_\beta\right) & = & A_{\alpha\beta} \sin^2 \Delta_{atm} \qquad (\alpha\ne\beta) \nonumber\\ {\cal P}_{LBL}\left(\nu_\alpha \rightarrow \nu_\alpha \right) & = & 1 - B_\alpha \sin^2\Delta_{atm} \ \ , \end{eqnarray} where ($\alpha,\beta = e, \mu, \tau$) \begin{eqnarray} A_{\alpha\beta} &=& 4 | U_{\alpha3}|^2 | U_{\beta3}|^2 \ \ ,\nonumber\\ B_\alpha &=& 4 | U_{\alpha3}|^2 \left( 1 - | U_{\alpha3}|^2 \right) \ \ . \end{eqnarray} Then eq.(\ref{eq:MNS2}) implies \begin{equation} \begin{tabular}{lll} $A_{e\mu} = s^2 c^2$ , \qquad & $ A_{e\tau} = s^2$ , \qquad & $A_{\mu\tau}= c^2$ ,\\ $B_e = 1 - c^4$ , \qquad & $ B_\mu = 1 -s^4$ , \qquad & $ B_\tau = 1$ ,\\ \end{tabular} \end{equation} and using (\ref{eq:s}) we estimate that for appearance type $\nu_e \rightleftharpoons \nu_\mu$ and $\nu_e \rightleftharpoons \nu_\tau$ LBL experiments \begin{equation} \sin^2 2\vartheta_{a} \approx 0.005 \label{eq:appear} \end{equation} whereas \begin{equation} \sin^2 2\vartheta_{d} \approx 0.01 \label{eq:disappear} \end{equation} for disappearance of $\nu_e$ and $\bar\nu_e$ neutrinos. These predictions can be tested in future high precision experiments. However the numbers are rather small. The estimation (\ref{eq:appear}) is about two times smaller than planned sensitivity of MINOS\cite{MINOS} and eq.(\ref{eq:disappear}) predicts disappearance of electron neutrinos at the level 20 times smaller than the present limit from CHOOZ\cite{CHOOZ}. Still these estimations are larger than those following from SMA MSW solutions for which $\sin^2 2\vartheta_{d}$ is driven to much smaller values by the recent data on day-night asymmetries for solar neutrinos\cite{day-night}. For solar neutrinos, c.f.\cite{BiGG} and references therein, \begin{equation} {\cal P}_\odot \left(\nu_e\rightarrow\nu_e\right) = \left( 1 - | U_{e3} |^2 \right) \left( 1 - \sin^2 2\vartheta_\odot\; \sin^2\Delta_\odot \right) + | U_{e3} |^4\ \ , \end{equation} where \begin{equation} \sin^2 2\vartheta_\odot = 4 | U_{e1} |^2\, | U_{e2} |^2\, / \, \left( 1 - | U_{e3} |^2 \right)^2 \ \ . \end{equation} For the mixing matrix (\ref{eq:MNS2}) \begin{equation} {\cal P}_\odot = \left( 1- s^2\right) \left( 1 - \sin^2 2\vartheta_\odot\; \sin^2\Delta_\odot \right) + {\cal O}(s^4)\ \ , \end{equation} with \begin{equation} \sin^2 2\vartheta_\odot \approx 1 - s^2 \approx 0.99 \quad . \label{eq:thetasun} \end{equation} It is evident that the simple picture proposed in this paper can be falsified when $\sin^2 2\vartheta_\odot$ derived from solar neutrino data is significantly below 1. In Ref.\cite{BKS} the best fit to vacuum oscillations is \begin{equation} \Delta m^2_\odot = 6.5 \times 10^{-11}\ \mbox{eV}^2, \qquad \sin^2 2\vartheta_\odot = 0.75 \label{eq:BKS} \end{equation} which seems to be in conflict with the estimation in (\ref{eq:thetasun}). In Ref.\cite{BW99a}, see also \cite{BFLi}, a new preliminary data from SuperKamiokande \cite {Suzuki99} are included into analysis leading to acceptable solutions in three regions (called $A$, $C$ and $D$) of parameter space \begin{equation} \matrix{ C: \qquad & \Delta m^2_\odot = 4.4 \times 10^{-10}\ \mbox{eV}^2, \qquad & \sin^2 2\vartheta_\odot = 0.93 , \qquad & gof = 14\% \cr D: \qquad & \Delta m^2_\odot = 6.4 \times 10^{-10}\ \mbox{eV}^2, \qquad & \sin^2 2\vartheta_\odot = 1.00 , \qquad & gof = 8\% \cr A: \qquad & \Delta m^2_\odot = 6.5 \times 10^{-11}\ \mbox{eV}^2, \qquad & \sin^2 2\vartheta_\odot = 0.70 , \qquad & gof = 6\% \cr} \label{eq:BW99a} \end{equation} In eq.(\ref{eq:BW99a}) the parameter $gof$ (goodness-of-fit) is the probability that a random repeat of the given experiment would observe a greater $\chi^2$, assuming the model is correct. The region $A$ is the same as selected by the fits of Ref.\cite{BKS}. Clearly the analysis of Ref.\cite{BW99a} gives results in better agreement with eq.(\ref{eq:thetasun}). It is also seen that new data on solar neutrinos will seriously test the estimations presented in this article. \section{Neutrinoless double beta decay} Let us discuss now predictions for neutrinoless double beta decays ($0\nu2\beta$)\footnote{A model independent analysis is presented in a recent preprint\cite{BW99b}.}. Probabilities of such transitions depend on a mass parameter $B$ which is equal to the absolute value of the element $\left( N_A \right)_{11}$ of the matrix \begin{equation} N_A = U_{MNS}\, M_{L,A}\, {U_{MNS}}^T \quad . \end{equation} $M_{L,A}$ (for $A = I, II, III, IV$) denotes a diagonal matrix of Majorana masses for the three light neutrinos: \begin{equation} M_{L,A} = m\; diag(\lambda_1,\lambda_2,\lambda_3) \quad , \end{equation} where $|\lambda_i| \le 1 + {\cal O}(\epsilon,\xi)$; $\epsilon$ and $\xi$ are small parameters to be discussed in the following section. In our convention the eigenvalues $\lambda_i$ are ordered in such a way that \begin{equation} \Delta m^2_\odot/ m^2 = | \lambda_1^2 - \lambda_2^2 | \end{equation} and \begin{equation} \Delta m^2_{atm}/ m^2 \approx | \lambda_1^2 - \lambda_3^2 | \approx | \lambda_2^2 - \lambda_3^2 | \quad . \end{equation} For bimaximal mixing $\nu_e$ is an equal mixture of the mass eigenstates 1 and 2. Considering $\Delta m^2_\odot /\Delta m^2_{atm}$ as a small perturbation one obtains $|\lambda_1| = |\lambda_2|$. Then, up to irrelevant sign redefinitions there are only four mass patterns consistent with the data on $0\nu 2\beta$: \begin{enumerate} \item[I)] $\qquad m = \sqrt{\Delta m^2_{atm}}$\ , \quad $\lambda_1 = \lambda_2 = 0$, $\lambda_3 = 1$ \item[II)] $\qquad m = \sqrt{\Delta m^2_{atm}}$\ , \quad $\lambda_1 = -1$, $\lambda_2 = 1$, $\lambda_3 = 0$ \item[III)] $\qquad m = \sqrt{\Delta m^2_{atm}}$\ , \quad $\lambda_1 = \lambda_2 = 1$, $\lambda_3 = 0$ \item[IV)] $\qquad m = {\cal O}(1\ \mbox{eV}) \gg \sqrt{\Delta m^2_{atm}}$\ , \quad $\lambda_1 = -1$, $\lambda_2 = 1$, $\lambda_3 = \eta$ where $\eta = \pm 1$ . \end{enumerate} In case $IV$ $\Delta m^2_{atm}/m^2$ is considered as a small perturbation\footnote{The case $\Delta m^2_{atm}/m^2 ={\cal O}(1)$ requires much more tuning of parameters than the case $IV$. We do not discuss this case in the present article because it is not clear if a reasonable model of this kind exists. Furthermore an original motivation for degenerate neutrino masses is an appreciable neutrino contribution to dark matter in the Universe whereas this contribution is small if $\Delta m^2_{atm}/m^2$ is not small.}. The states 1 and 2 have opposite CP parities (i.e. $\lambda_1$ and $\lambda_2$ have opposite signs, see e.g.\cite{BiGG}) due to the experimental bound\cite{Baudis}: \begin{equation} B < 0.2\ \mbox{eV} \end{equation} which is violated for $m > 0.2\ eV$ and $\lambda_1 = \lambda_2 = 1$. The argument is analogous to the case $III$ which is considered in the following, c.f. eq.(\ref{eq:NIII11}). \par\noindent If corrections to bimaximal mixing are neglected the following mass matrices are obtained \begin{eqnarray} {\bar N_I} &=& {\textstyle {1\over2} } \sqrt{\Delta m^2_{atm}} \left( \matrix{0 & 0 & 0 \cr 0 & 1 & 1 \cr 0 & 1 & 1 \cr } \right) \quad , \label{eq:textureI}\\ {\bar N_{II}} &=& {\textstyle {1\over\sqrt{2}}} \sqrt{\Delta m^2_{atm}} \left( \matrix{\ 0 & 1 & -1 \cr \ 1 & 0 & \ 0 \cr -1 & 0 & \ 0 \cr } \right) \quad , \label{eq:textureII}\\ {\bar N_{III}} &=& {\textstyle {1\over2} } \sqrt{\Delta m^2_{atm}} \left( \matrix{2 & \ 0 & \ 0 \cr 0 & \ 1 & -1 \cr 0 & -1 & \ 1 \cr } \right) \quad , \label{eq:textureIII}\\ {\bar N_{IV} } &=& {\textstyle {1\over2} }\, m \, \left( \matrix{\ 0 & \sqrt{2} & -\sqrt{2} \cr \ \sqrt{2} & \eta & \ \eta \cr -\sqrt{2} & \eta & \ \eta \cr} \right)\quad , \label{eq:textureIV} \end{eqnarray} where for $ A = I, II, III, IV$: \begin{equation} \bar N_{A} = U_{bm}\; M_{L,A} \; {U_{bm}}^T \quad . \end{equation} In the following section we consider also matrices \begin{equation} \widetilde N_{A} = U_{+}\; M_{L,A} \; {U_{+}}^T \end{equation} which include deviations from the ideal bimaximal mixing due to the sector $I_3 = + 1/2$. The textures (\ref{eq:textureI})--(\ref{eq:textureIV}) have been discussed in the literature\cite{JS98}--\cite{FGN}. It is seen that only $\bar N_{III}$ leads to non-zero rates of $0\nu2\beta$ decays. In this case deviations from bimaximal mixing can be neglected and one obtains the following prediction \begin{equation} | N_{III} |_{11} \approx \sqrt{\Delta m^2_{atm}} = ( 6 \pm 2 ) \times 10^{-2}\ \mbox{eV} \quad . \label{eq:NIII11} \end{equation} In the three other cases the role of $U_{-}$ is essential. Keeping only leading terms one obtains \begin{eqnarray} | N_{I} |_{11} \approx& {\textstyle{1\over2}}\; s^2\; \sqrt{\Delta m^2_{atm}} &= ( 1.5 \pm 0.5 ) \times 10^{-4}\ \mbox{eV} \quad , \label{eq:NI11}\\ | N_{II} |_{11} \approx& \sqrt{2}\, |s|\, \sqrt{\Delta m^2_{atm}} &= ( 6 \pm 2 ) \times 10^{-3}\ \mbox{eV} \quad , \label{eq:NII11}\\ | N_{IV} |_{11} \approx& \sqrt{2}\, |s|\, m &\approx 0.1\, m \quad . \label{eq:NIV11} \end{eqnarray} Evidently the cases $III$ and $IV$ can be confirmed or ruled out by next generation experiments which may be sensitive to $B$ as low as 0.01 eV\cite{K-KHH}. Even the present limit\cite{Baudis} implies that for degenerate masses $m$ must be smaller than 2~eV. \section{Patterns of neutrino masses} In this section effects are considered due to non-zero value of the ratio $\Delta m^2_\odot /\Delta m^2_{atm}$ which were neglected in the discussion of the preceding section. We present simple seesaw models which show that these corrections are quite small, so the estimations (\ref{eq:NIII11})--(\ref{eq:NIV11}) are not much affected. In the following we consider the four cases $I$--$IV$ corresponding to different neutrino mass patterns. \subsection{Case I} A class of seesaw models corresponding to case $I$ has been described in\cite{JS98}. In these models the third mass eigenstate is much heavier than the other two whose masses are fairly close. In the following this mass pattern ($m_1\approx m_2 \ll m_3$) is called {\em semi-hierarchical}. In terms of a small parameter \begin{equation} |\epsilon| = \left( {\Delta m^2_\odot\over 2\Delta m^2_{atm}} \right)^{1/3} \approx 0.004 \end{equation} the eigenvalues $\lambda_i$ are given by the expansions: \begin{eqnarray} \lambda_1 &=& \ \epsilon - \epsilon^2/2 + \dots \nonumber\\ \lambda_2 &=& -\epsilon - \epsilon^2/2 + \dots \nonumber\\ \lambda_3 &=& 1 + \epsilon^2 + \dots \label{eq:semi-hierarch} \end{eqnarray} For semi-hierarchical mass pattern $\nu_e$ is a linear combination of the two lighter mass eigenstates of opposite CP parities. The element $(\widetilde N_{I} )_{11}$ of the matrix \begin{equation} \widetilde N_{I} = U_{+}\; M_{L,I} \; {U_{+}}^T \end{equation} is strongly suppressed and the main contribution to $0\nu 2\beta$ transitions originates from the mixing in the sector $I_3 = - 1/2$. The estimation of $| N_{I} |_{11}$ given in eq.(\ref{eq:NI11}) is not affected as can be seen from the explicit form of the matrix $U_{+}:$\footnote{ In notation of Ref.\cite{JS98} $r = 2^{3/2}\epsilon$. The columns of the matrix $U_{+}$ in eq.(\ref{eq:U+1/2a}) are normalized eigenvectors $v_1$, $- v_2$ and $v_3$, c.f. eq.(29) in \cite{JS98}. We have corrected a misprint in the second element of the vector $v_2$ in \cite{JS98}.} \begin{equation} U_{+} = \left( \matrix{ ( 1 + \epsilon/2 )/\sqrt{2} & ( 1 - \epsilon/2 )/\sqrt{2} & \epsilon\cr - 1/2 - \epsilon & 1/2 - \epsilon & 1/\sqrt{2} \cr 1/2 & - 1/2 & 1/\sqrt{2} \cr } \right) + {\cal O}(\epsilon^2) = U_{bm} + {\cal O}(\epsilon) \label{eq:U+1/2a} \end{equation} and \begin{equation} \widetilde N_I = {\textstyle {1\over 2}} \left( \matrix{ 0 & 0 & 2^{3/2}\epsilon \cr 0 & 1 & 1 \cr 2^{3/2}\epsilon & 1 & 1 \cr } \right) + {\cal O}(\epsilon^2) \quad . \label{eq:tildeNI} \end{equation} In case $I$ a strongly ordered {\em hierarchical} mass pattern ($m_1 \ll m_2 \ll m_3$) is obtained if e.g. \begin{equation} \lambda_1 = \epsilon^2 + \dots \ , \qquad \lambda_2 = \epsilon + \dots \ , \qquad \lambda_2 = 1 + \dots \ , \end{equation} with \begin{equation} |\epsilon| = \sqrt{\Delta m^2_\odot /\Delta m^2_{atm}} \approx 3.7 \times 10^{-4} \ \ . \end{equation} The magnitude of $\epsilon$ is an order of magnitude smaller than for semi-hierarchical spectrum. There is no suppression of $\left( N_{I} \right)_{11}$ because the lighter states have the same CP parity but the parameter $\epsilon$ is small. Seesaw models leading to hierarchical spectrum may require some fine tuning. Thus we believe that hierarchical mass pattern is less attractive from theoretical point of view. In spite of theoretical prejudices we note that for both patterns predictions for oscillations and $0\nu 2\beta$ decays are similar. \subsection{Case II} Corresponding seesaw models can be easily derived. An example is (in notation of \cite{JS98}): \begin{equation} {\bf m_{_D}} \sim \left( \matrix{ x^2 y & 0 & 0 \cr 0 & x & -x \cr 0 & x^2 & 1 \cr} \right) \qquad {\rm and} \qquad {\bf M_{_R}^{-1}} \sim \left( \matrix{ a_{11} & a_{12} & 0 \cr a_{12} & 0 & 0 \cr 0 & 0 & 0 \cr} \right)\quad , \end{equation} where ${\bf m_{_D}}$ is the Dirac mass matrix and ${\bf M_{_R}^{-1}}$ denotes inverse of the Majorana mass matrix for the righthanded neutrinos\footnote{ This form of ${\bf M_{_R}^{-1}}$ means that either one of the righthanded neutrinos is much heavier and decoupled or there are only two heavy neutrinos in the particle spectrum and their couplings to the lefthanded neutrinos are given by the matrix $$ {\bf m_{_D}} \sim \left( \matrix{ xy & 0 & 0 \cr 0 & 1 & -1 \cr} \right) .$$ }. Then \begin{equation} \widetilde N_{II} = U_{+}\; M_{L,II} \; {U_{+}}^T = {\bf m_{_D}}^T\, {\bf M_{_R}^{-1}}\, {\bf m_{_D}} = \sqrt{\Delta m^2_{atm}}\, \left( \matrix{ 2 \epsilon & 1/\sqrt{2} & -1/\sqrt{2} \cr 1/\sqrt{2} & 0 & 0 \cr -1/\sqrt{2} & 0 & 0 \cr} \right)\quad , \end{equation} with the eigenvalues \begin{equation} \lambda_1 = -\sqrt{1 + \epsilon^2} + \epsilon\ , \qquad \lambda_2 = \ \sqrt{1 + \epsilon^2} + \epsilon\ , \qquad \lambda_3 =0 \end{equation} and \begin{equation} |\epsilon| = 2^{-3/2} xy |a_{11}/a_{12}| = \Delta m^2_\odot / (4 \Delta m^2_{atm}) = 3.4 \times 10^{-8}\quad . \end{equation} Evidently corrections of order $\epsilon$ to $U_{MNS}$ are negligibly small. The spectrum of light neutrinos contains two heavier states. The electron neutrino is a nearly maximal mixture of these states with a tiny admixture of the third state which is massless. We call this spectrum {\em semi-degenerate} ($m_1 \approx m_2 \gg m_3$). The ratio $|a_{11}/a_{12}|$ must be very small which implies that the two heavy Majorana neutrinos form a pseudo-Dirac system\cite{BHS98}. In our opinion the semi-degenerate mass pattern is more attractive for solutions of the solar neutrino problem with larger values of $\Delta m^2_\odot / \Delta m^2_{atm}$~\cite{BGG}, see also discussion in \cite{FGN}. \subsection{Case III} As in case $II$ the parameter $\epsilon$ is tiny and the related corrections are negligible. \subsection{Case IV} For {\em degenerate} neutrino masses\cite{Vissani97,GG98} one can choose the eigenvalues $\lambda_i$ in the following way: \begin{equation} \lambda_1 = -1 + \epsilon\ , \qquad \lambda_2 = 1 + \epsilon\ , \qquad \lambda_3 = \eta + \xi \ , \end{equation} with \begin{equation} |\epsilon| = \Delta m^2_\odot / (4 m^2) \qquad {\rm and} \qquad |\xi| = \Delta m^2_{atm} /(2 m^2) \quad . \end{equation} Corrections to eq.(\ref{eq:NIV11}) are small for $|\xi| \ll |s|$. If the case of degenerate neutrino masses is a reasonable option for vacuum oscillations remains an open question. On one side dynamical models have been recently found\cite{BHKR} for the degenerate mass pattern. On the other side the parameter $\epsilon$ is extremely small and the problem of stability may be very serious, see \cite{ELola}. It is also not clear if neutrino masses in the range of 1~eV are really needed by cosmology, see e.g. \cite{Turner}. Fortunately it is exactly this mass range and mass pattern for which our estimations give numbers close to the present experimental limit for $0\nu 2\beta$ decays. Therefore one may hope that a definite answer will come from the experiment in not very distant future. \section{Summary} Neutrino oscillations and Majorana masses are discussed assuming bimaximal mixing of leptons. In this scheme the atmospheric neutrino anomaly~\cite{SuperK98a} is described as a result of $\nu_\mu\to \nu_\tau$ oscillations and deficits of solar neutrinos~\cite{solarEx} as vacuum oscillations of $\nu_e$. A simple hypothesis is proposed that deviations from bimaximal mixing are dominated by a rotation in the electron--muon plane of the charged lepton sector $I_3 = - 1/2$. This hypothesis is in agreement with the present data and can be tested by future precision data on solar neutrinos. Predictions for oscillations in long baseline experiments are given. There are four classes of neutrino mass patterns leading to quite different predictions for the rates of neutrinoless double beta decays. \vskip0.5cm\par\noindent {\Large\bf Acknowledgements} \vskip0.5cm\par\noindent M.J. thanks Wolfgang Hollik, Hans K\"uhn, Thomas Mannel and Marek Zra{\l}ek for useful comments and remarks. Y.S. is grateful to K. Inoue for discussions. M.J. would like to acknowledge a support from BMBF (FRG) under Project Nr. POL-239-96, and a very stimulating atmosphere and warm hospitality during his stay in the Institut f\"ur Theoretische Teilchenphysik, Universit\"at Karlsruhe where a large part of this work was done. \newpage
\section{\@startsection{section}{1}{\z@}% {-3.5ex plus -1ex minus -.5ex}{1.5ex plus.3ex}{\bf }} \def\subsection{\@startsection{subsection}{1}{\z@}% {-3.5ex plus-1ex minus-.5ex}{1.5ex plus.3ex}{\bf }} \begin{document} \hfill\parbox{4.77cm}{\Large\centering Annalen\\der Physik\\[-.2\baselineskip] {\small \underline{\copyright\ Johann Ambrosius Barth 1998}}} \vspace{.75cm}\newline{\Large\bf Numerical tests of conjectures of conformal field theory for three-dimensional systems }\vspace{.4cm}\newline{\bf Martin Weigel and Wolfhard Janke }\vspace{.4cm}\newline\small Institut f\"{u}r Theoretische Physik, Universit\"{a}t Leipzig, D-04109 Leipzig, Germany, and\\ Institut f\"{u}r Physik, Johannes Gutenberg-Universit\"{a}t Mainz, D-55099 Mainz, Germany\\ {\tt <EMAIL>, <EMAIL>} \vspace{.2cm}\newline Received 6 October 1998, accepted 8 October 1998 by U. Eckern \vspace{.4cm}\newline\begin{minipage}[h]{\textwidth}\baselineskip=10pt {\bf Abstract.} The concept of conformal field theory provides a general classification of statistical systems on two-dimensional geometries at the point of a continuous phase transition. Considering the finite-size scaling of certain special observables, one thus obtains not only the critical exponents but even the corresponding {\em amplitudes} of the divergences analytically. A first numerical analysis brought up the question whether analogous results can be obtained for those systems on three-dimensional manifolds. Using Monte Carlo simulations based on the Wolff single-cluster update algorithm we investigate the scaling properties of O($n$) symmetric classical spin models on a {\em three-dimensional}, hyper-cylindrical geometry with a toroidal cross-section considering both periodic and antiperiodic boundary conditions. Studying the correlation lengths of the Ising, the XY, and the Heisenberg model, we find strong evidence for a scaling relation analogous to the two-dimensional case, but in contrast here for the systems with {\em antiperiodic} boundary conditions. \end{minipage}\vspace{.4cm} \newline {\bf Keywords:} Spin models; Finite-size scaling; Universal amplitudes; Conformal field theory \newline\vspace{.2cm} \normalsize \section{Introduction} Statistical mechanical systems at a critical point are essentially characterized by a loss of length scales: as the correlation length diverges, the system becomes a self-similar random fractal. Augmenting this symmetry with translational and rotational invariance in the continuum limit establishes the so-called {\em conformal} invariance of the system. As the 2D conformal group is of infinite dimension, exploiting this feature allows a complete classification of models of statistical mechanics according to their operator content in two dimensions \cite{CardyBuch,HenkelBuch}. This includes a special class of formally model independent relations, like finite-size scaling (FSS) laws; in particular, for the 2D strip geometry $S^1\times{\rm I\!R}$ with periodic boundary conditions Cardy \cite{Cardy84a} has shown, that for any primary, i.e. conformally covariant, operator of a model showing critical behavior the corresponding correlation length scales as: \begin{equation} \xi_i=\frac{A}{x_i}L, \label{amplit} \end{equation} where $L$ denotes the circumference of the cylinder, $x_i$ is the scaling dimension of the considered operator, a combination of the classical critical exponents, and $A=1/2\pi$ in the 2D case. For 3D systems, however, the situation is different. First, the concept of a primary operator becomes at least problematic (it might possibly be established in terms of the operator product expansion (OPE) \cite{CardyPrivate}). Secondly, numerically feasible and more widely applicable geometries like that of a column $S^1\times S^1\times{\rm I\!R}$ are not conformally related to flat spaces like in the above mentioned 2D case, which is a essential ingredient of the derivation of relation (\ref{amplit}). A transfer matrix calculation by Henkel \cite{Henkel86,Henkel87} for the $S=\frac{1}{2}$ Ising model on the column geometry gave for the ratio of the correlation lengths of the magnetization and energy densities the values $\xi_\sigma/\xi_\epsilon=3.62(7)$ and $2.76(4)$ for periodic and antiperiodic boundary conditions, respectively. Comparing this with the ratio of scaling dimensions of $x_\epsilon/x_\sigma=2.7326(16)$ \cite{diplom} results in the astonishing, theoretically obscure, conjecture that a relation of the form (\ref{amplit}) can be re-established in the 3D case, when applying {\em antiperiodic} (apbc) instead of periodic (pbc) boundary conditions along the torus directions, which later on could be affirmed by a Metropolis Monte Carlo (MC) simulation by Weston \cite{Weston}. If this result could be established analytically, it would be one of the rare rigorous statements for non-trivial 3D systems. As simulational data were available up to this point only for the single special case of the Ising model, we thought it rewarding to analyze some further models, so possibly establishing this conjecture at an empirical level, which should constitute a motivation and basis for further analytical studies. \section{Models} In generalization to the Ising case we restrict ourselves to the class of O($n$) symmetric classical spin models with Hamiltonian \begin{equation} \label{Hamilton} {\cal H} = -J \sum_{<ij>} {\bf s}_i\cdot{\bf s}_j,\;\;{\bf s}_i \in S^{n-1}, \end{equation} assuming nearest-neighbor, ferromagnetic ($J>0$) interactions. The simulations were done for a discrete sc lattice with dimensions $(L_x,L_y,L_z)$, choosing $L_x=L_y$ and $\xi/L_z\ll 1$, therewith approximately modelling the column geometry $S^1\times S^1\times{\rm I\!R}$. \begin{figure}[tb] \begin{picture}(250,200) \put(0, 0){\special{psfile=figure1.ps angle=-90 hscale=45 vscale=45 hoffset=8 voffset=240}} \end{picture} \caption{ FSS plot for the spin correlation length $\xi_\sigma(L_x)$ of the 3D Ising model with antiperiodic boundary conditions. The solid line represents a least-square fit according to Eq.\ (\ref{fitform}).} \label{fig1} \end{figure} \section{Simulation and data analysis} The MC simulations were done using the Wolff single-cluster updating scheme \cite{Wolff89}, as it is known to be more efficient than the Swendsen-Wang \cite{Swendsen} update in three dimensions \cite{WJChem}. In order to be able to perform simulations for both, periodic and antiperiodic boundary conditions, the Wolff update had to be adapted to the latter case: this was achieved by exploiting the fact that antiperiodic bc are equivalent to the insertion of a seam of anti-ferromagnetic bonds along the boundary in the case of nearest-neighbor interactions. To test for a relation according to (\ref{amplit}) we had to measure at least two different correlation lengths of the systems under consideration. Following Henkel and in an analogy to the 2D Ising case, where the only non-trivial primary operators are the densities of magnetization and energy, we recorded the correlation functions of these two operators: \begin{equation} \begin{array}{rcl} G_{\sigma}^c({\bf x}_1,{\bf x}_2) & = & \langle{\bf s}({\bf x}_1)\cdot{\bf s}({\bf x}_2)\rangle-\langle{\bf s}\rangle\langle{\bf s}\rangle, \\ G_{\epsilon}^c({\bf x}_1,{\bf x}_2) & = & \langle\epsilon({\bf x}_1)\,\epsilon({\bf x}_2)\rangle-\langle\epsilon\rangle\langle\epsilon\rangle. \\ \end{array} \label{conncorr} \end{equation} Variance-reduction of the estimators for these observables can be achieved in a first step by the trivial average over values with $({\bf x}_1-{\bf x}_2)\parallel \hat{e}_z$ and $i\equiv|{\bf x}_1-{\bf x}_2|=\mbox{const}$; they can be further improved by applying a zero momentum mode projection, i.e., by summing up the values for the densities in the layers $z=\mbox{const}$ before correlating them \cite{WJ93}. \begin{figure}[tb] \begin{picture}(250,200) \put(0, 0){\special{psfile=figure2.ps angle=-90 hscale=45 vscale=45 hoffset=8 voffset=240}} \end{picture} \caption{ Scaling of the amplitudes $\xi_\sigma/L_x$ of the 3D Ising model with antiperiodic boundary conditions.} \label{fig2} \end{figure} \begin{table}[tb] \caption{Finite-size scaling amplitudes of the correlation lengths of the Ising, XY, and Heisenberg models on the $T^2\times{\rm I\!R}$ geometry. \label{tab1}} \begin{center} \begin{tabular}{||clll||} \hline model & & \multicolumn{1}{c}{pbc} & \multicolumn{1}{c||}{apbc} \\ \hline & $A_\sigma$ & 0.8183(32) & 0.23694(80) \\ & $A_\epsilon$ & 0.2232(16) & 0.08661(31) \\ \raisebox{1ex}[-1ex]{Ising} & $A_\sigma/A_\epsilon$ & 3.666(30) & 2.736(13) \\ & $x_\epsilon/x_\sigma$ & \multicolumn{2}{c||}{2.7326(16)} \\ \hline & $A_\sigma$ & 0.75409(59) & 0.24113(57) \\ & $A_\epsilon$ & 0.1899(15) & 0.0823(13) \\ \raisebox{1ex}[-1ex]{XY} & $A_\sigma/A_\epsilon$ & 3.971(32) & 2.930(47) \\ & $x_\epsilon/x_\sigma$ & \multicolumn{2}{c||}{2.923(7)} \\ \hline & $A_\sigma$ & 0.72068(34) & 0.24462(51) \\ & $A_\epsilon$ & 0.16966(36) & 0.0793(20) \\ \raisebox{1ex}[-1ex]{Heisenberg} & $A_\sigma/A_\epsilon$ & 4.2478(92) & 3.085(78) \\ & $x_\epsilon/x_\sigma$ & \multicolumn{2}{c||}{3.091(8)} \\ \hline \end{tabular} \end{center} \end{table} For extracting the correlation lengths from (\ref{conncorr}) one can cancel out deviations that arise from inaccuracies in the determination of the disconnected parts of the correlation functions and eliminate the need for a correct normalization of the estimates by considering the following set of estimators: \begin{equation} \hat{\xi}_i=\Delta{\left[\ln\frac{\hat{G}^{c,\parallel}(i)-\hat{G}^{c,\parallel}(i-\Delta)} {\hat{G}^{c,\parallel}(i+\Delta)-\hat{G}^{c,\parallel}(i)}\right]}^{-1}, \label{diffmethoddelta} \end{equation} where $\Delta\ge 1$ should usually be chosen so that a constant drop of $G(i)$ between the pairs of considered points is guaranteed. Variances and cross-correlations of the $\hat{\xi}(i)$ were estimated using a combined binning \cite{Flyvbjerg} and jackknifing \cite{Efron} technique. In a process of statistical optimization, resulting in the leaving out of the corrupt estimates $\hat{\xi}(i)$ for distances in the regions $i<\Delta$ and $i>L_z/2$, one ends up with a final value for the correlation lengths $\xi_\sigma$ and $\xi_\epsilon$ of the considered system. All simulations were done at inverse temperatures, which were either highly precise single estimates of the inverse critical temperature of the bulk model or weighted means of several such estimates \cite{cond,ours}, the influence of uncertainties in these values being checked via a temperature reweighting technique. \section{Results} The cumulated estimates for the correlation lengths $\hat{\xi}(L_x)$ for the different system sizes exhibit an almost perfect linear scaling behavior as shown in Fig. \ref{fig1} for the Ising model and antiperiodic bc. For all models we analyzed system sizes between $L_x=4$ and $30$ and volumes up to about $3\cdot 10^5$ spins. The plot of the amplitudes $\hat{\xi}/L_x$ in Fig. \ref{fig2}, however, reveals a clear resolution of corrections to scaling, however. Therefore fits to a law including corrections of the form \begin{equation} \xi(L_x)=AL_x+BL_x^{\alpha} \label{fitform} \end{equation} were done to arrive at estimates for the leading order scaling amplitudes $A_\sigma$ and $A_\epsilon$. The final results for these leading amplitudes are summarized in Table \ref{tab1}. The values for the scaling dimensions shown for comparison are once again weighted literature means \cite{diplom}. \section{Conclusions} The clear conclusion of these results for all three models under consideration is that, while for the generic case of periodic bc the ratios of the amplitudes and the scaling dimensions differ by at least about thirty sigma, in the case of {\em antiperiodic} bc both ratios agree to a very high level of precision, thus giving the conjecture of a law equivalent to (\ref{amplit}) for 3D models enough backing to think seriously about a theoretical justification. So, for the $n=1,2,3$ representatives of the class of O($n$) spin models one can state that the finite-size scaling amplitude ratios of the the correlation lengths of the magnetization and energy densities is determined by the corresponding scaling dimensions and thus universal; in connection with additional results for the case $n=10$ \cite{ours} and an analytical results for the spherical model \cite{Henkel88}, it seems reasonable to assume that such a relation holds for the whole class of O($n$) spin models. Note, however, that the amplitude $A$ in Eq. (\ref{amplit}), which was $1/2\pi=\mbox{const}$ in the 2D case, does now depend on the model under consideration, i.e., the dimension $n$ of the order parameter \cite{cond}. \vspace{0.6cm}\newline{\small We thank K. Binder for his constant and generous support. We are grateful to J. Cardy and M. Henkel for helpful discussions on the theoretical background. W.J. gratefully acknowledges support from the Deutsche Forschungsgemeinschaft through a Heisenberg Fellowship. }
\section{INTRODUCTION} \label{sec:intro} In the standard picture of cosmology, structure originated from small-amplitude density fluctuations that were amplified by gravitational instability. These initial fluctuations are assumed to have a Gaussian probability distribution, fully characterized by their power spectrum (PS). On large scales, the fluctuations are expected to be linear even at late times, still characterized by the initial PS. Thus, the PS is a very useful statistics for large scale-structure. We focus on the PS rather than the correlation function ({\it e.g.}, G\'orski {\it et al.\ } 1989) because the PS distinguishes more clearly between the processes that affect structure formation on different scales. It also has the advantage of being less sensitive to assumptions regarding the mean density. The PS has been estimated from several redshift surveys of galaxies (see reviews by Strauss \& Willick 1995; Strauss 1998). However, the distribution of galaxies does not necessarily provide a direct measurement of the underlying {\it mass} distribution; the PS estimated from redshift surveys is contaminated by unknown ``galaxy biasing". Additional complications arise from redshift distortions, triple-value zones and nonlinearities of the density field. This also complicates a direct comparison of the correlation function derived from galaxy density fluctuations with similar quantities derived from peculiar velocity measurements. Therefore, it is advantageous to estimate the mass PS directly from purely dynamical data. Another advantage of velocity over density data is that they probe the density field on scales larger than the sample itself, and that they are subject to weaker nonlinear effects. It is therefore easier to obtain an approximation for the initial PS from the current velocity PS than from the current density PS. Direct estimation of the PS from reconstructed velocity or density fields is complicated by the need to correct for the effects of large noise, smoothing, and finite and nonuniform sampling ({\it e.g.}, Kolatt \& Dekel 1997). On the other hand, the likelihood analysis of peculiar velocities, such as the one applied here, provides an appealing method for estimating the mass PS since it is a straightforward statistic acting on the `raw' data, without the need for processing such as binning, smoothing, or applying a full POTENT reconstruction (Dekel, Bertschinger \& Faber 1990). It takes into account the measurement errors and finite discrete sampling, and it utilizes much of the information content of the data. The simplifying assumptions made in our main analysis are that the peculiar velocities are drawn from a Gaussian random field, that the velocity correlations can be derived from the density PS using linear theory, and that the errors in the measurements are Gaussian. Other limitations of the method are the need to assume some parametric functional form for the PS, with a possible sensitivity of the results to the choice of this model, and the fact that the likelihood analysis provides only relative likelihood of the different models, not an absolute goodness-of-fit. PS estimates using likelihood analysis (Zaroubi {\it et al.\ } 1997) has been obtained from the ``Mark III'' catalog of peculiar velocities (Willick {\it et al.\ } 1997a), yielding relatively high values for the PS, in agreement with the direct estimates from the ``POTENT'' reconstruction (Kolatt \& Dekel 1997). This result is still associated with large uncertainties because the sampling of the data is sparse and nonuniform, because the merging of data from several sources is nontrivial, and because the distance errors in peculiar-velocity data are relatively large. Furthermore, the uncertainty in the assumed distance errors always propagates into an uncertainty in the resultant PS because the errors add in quadrature to the PS. It is therefore important to analyze new data of certain improved qualities and to pay special attention to the error estimates. The data analyzed in the present paper are based on the new SFI catalog of peculiar velocities of galaxies (Haynes {\it et al.\ } 1999, Wegner {\it et al.\ } 1999), containing about 1300 field spiral galaxies with Tully-Fisher (TF) distances. Most of the measurements in the SFI catalog are new. Data taken from the literature which are included in the catalog, mostly those by Mathewson, Ford \& Buchhorn (1992), have been recalibrated to match the new observations both for magnitude and line width scale. This procedure should minimize the effects of combining different datasets, effects of significant concern in Mark~III (e.g., Willick \& Strauss 1998). The SFI catalog, though sparser than Mark III in certain places, covers more uniformly the volume out to $70\,{\rm h^{-1}Mpc}$. The distances in the SFI catalog have been estimated using a linear TF relation derived from a matching cluster sample (Giovanelli {\it et al.\ } 1997a, 1997b; SCI). Possible deviations from the standard, linear TF relation were ignored since no clear evidence for such deviations was detected in the data. In addition, the sensitivity to such an effect is small because of the the selection criteria of the SFI catalog. The crucial issue of error estimate is addressed in two ways. First, the fact that the SFI field sample is matched by the SCI cluster sample of similar size allows a careful investigation of the observational and internal scatter of the TF distances which provides a good a priori estimate of the errors. These errors are adjusted for an assumed difference in the scatter between field and cluster galaxies. An additional adjustment of the scatter is due to our bias-correction procedure. A second and independent approach to estimate the errors is to include them as an extra parameter in the likelihood analysis so that it also determines the maximum-likelihood values for the errors. In that approach, we use a parametric model for the errors which builds upon the original estimates of width-dependent errors. We address here the mass-density power spectrum as derived from peculiar-velocity data, with or without Cosmic Microwave Background (CMB) fluctuation data, but independent of the distribution of galaxies in redshift space. We thus determine the quantity $P(k) \Omega^{1.2}$ (where $P(k)$ is the density power spectrum and $\Omega$ is the cosmological density parameter), while we are free of assumptions regarding the ``biasing" relation between galaxies and mass. We can therefore measure a purely dynamical parameter such as $\tilde\sigma_8\equiv\sigma_8 \Omega^{0.6}$ (where $\sigma_8$ is the rms mass-density fluctuation in top-hat spheres of radius $8\,{\rm h^{-1}Mpc}$). When assuming a priori a parametric functional form for the mass PS, {\it e.g.}, based on a generic CDM model, we can in fact determine a combination of dynamical parameters such as $\Omega$ and the power index $n$. Investigations involving galaxy redshift surveys commonly measure a different parameter that does involve galaxy biasing, $\beta\equiv\Omega^{0.6}/b$ (where $b$ is the biasing parameter). The parameters $\tilde\sigma_8$ and $\beta$ (at $8\,{\rm h^{-1}Mpc}$) are related via $\sigma_{8\rm g}$, referring to the rms fluctuation in the galaxy number density. A number of measurements of $\beta$ have been carried out so far, either based on redshift distortions of the IRAS 1.2 Jy redshift survey (Fisher {\it et al.\ } 1995) or based on comparisons of this redshift survey and the peculiar-velocity data. Most recent velocity-velocity comparisons found values for $\beta$ in the range of $0.5-0.7$ (Davis, Nusser \& Willick 1996; Willick {\it et al.\ } 1997b; da Costa {\it et al.\ } 1998; Kashlinsky 1998; Willick \& Strauss 1998), while density-density comparisons have lead to values as high as 0.9 (e.g. Sigad {\it et al.\ } 1998). A determination of $\tilde\sigma_8$ from the SFI data may help to clarify the situation. In \S~\ref{sec:data} we describe the data and our method for correcting Malmquist bias. In \S~\ref{sec:method} we present the method of analysis and the parametric models used as priors. The method is tested using mock catalogs in \S~\ref{sec:mock}. The estimated power spectra and the constraints on the cosmological parameters are presented in \S~\ref{sec:results}. The robustness of the results is addressed in \S~\ref{sec:robust}. We discuss our results and conclude in \S~\ref{sec:concl}. \section{DATA} \label{sec:data} \subsection{Sample and Distance Errors} \label{subsec:derrors} The SFI sample is based on a wide-angle survey of Sbc-Sc galaxies with I-band TF distances, covering declinations $\delta \ge -45^{o}$ and galactic latitudes $b \ge 10^{o}$. The galaxy selection criteria depend on redshift in order to ensure dense sampling at large distances; the catalog consists of three zones of different diameter limits and redshift limits. This data set was complemented south of $\delta=-45^{o}$ with galaxies drawn from the Mathewson {\it et al.\ } (1992) survey, carefully converted to the same system of magnitude and line-width, and with the same set of corrections and selection criteria applied to the whole sample (Giovanelli {\it et al.\ } 1997a, 1997b). The combined sample comprises of about 1300 field galaxies, extending out to $7500 \ {\rm km\,s^{-1}}$ in redshift, and quite isotropically covering the whole sky except of the Galactic zone of avoidance. Accurate estimation of the uncertainty $\Delta$ in the distance are important both for the bias correction (see \S~\ref{subsec:bias}) and for determination of the PS (see \S~\ref{sec:method}). The uncertainties are derived from the estimate of the scatter in the observed TF relation. We take advantage of the fact that the SFI sample is matched by a similar cluster sample (SCI, Giovanelli {\it et al.\ } 1997a, 1997b). The line-width dependent scatter of this cluster sample is well determined. Since SCI was observed using the same observational procedures as most galaxies in SFI, the distance estimates in both samples should suffer from similar observational uncertainties. However, it is less clear whether the intrinsic scatter of the TF relation is the same for the field and cluster samples. We have parameterized the total scatter in the SFI sample by using the SCI observed scatter and adding an additional intrinsic scatter for field galaxies in quadrature. Such a higher scatter for field galaxies has consistently been found by a number of authors (e.g. Bothun \& Mould 1987; Freudling, Martel \& Haynes 1991). We estimated the total scatter of SFI by taking advantage of the distance dependence of biases in the inferred distances. These biases for field galaxies are large at high distances and dominate the raw measured peculiar velocities. The exact behavior at large distance depends on the assumed amplitude of the scatter (see Freudling {\it et al.\ } 1995). With the aid of mock samples, the observed distance dependence of the average measured peculiar velocity was used to infer the intrinsic scatter for the SFI sample. The resulting errors are estimated to be in the range $15-20\%$, and increasing with decreasing line-width $w$. Following da Costa {\it et al.\ } (1996), a small fraction ($\sim 7\%$) of galaxies with small line-width (${\rm log} w \le 2.25$) has been discarded because of the unreliability of the TF relation and its scatter at such line-widths. A detailed account on the sample selection, error estimates, and the procedure of combining the two datasets can be found in Wegner {\it et al.\ } (1999) and Haynes {\it et al.\ } (1999). The SCI sample of $\sim 500$ galaxies within 24 clusters was used for calibrating the TF relation and in estimating the scatter properties, but the peculiar velocities of these clusters themselves are not used in this work as they require a different treatment. \subsection{Bias Correction} \label{subsec:bias} It is crucial to properly correct the data for systematic biases, such as those arising from the coupling between the random distance errors, the geometry of space and the inhomogeneities in the underlying distribution of galaxies, and certain aspects of the sample selection. Due to the complexity of the selection criteria and the TF distance errors in the SFI data, the bias correction could not be properly estimated using the standard simple analytic expression. In earlier papers of the SFI series, the bias was estimated using a numerical Monte-Carlo approach in which the selection criteria were mimicked in detail (Freudling {\it et al.\ } 1995). Here, we replace it with a simpler semi-analytic estimate of the bias, which incorporates the relevant selection criteria. The bias-correction method will be described in detail by Eldar {\it et al.\ } (1998). Here we mention only the basic features of the method. Given a galaxy with a TF inferred distance $d$ and a line-width $\eta={\rm log} w-2.5$, the Malmquist-corrected distance is adopted to be the conditional expectation value of the true distance, $r$, \begin{equation} E(r|d, \eta) = {\int _0^\infty \,dr\ r\ P(r,d,\eta)\over \int _0^\infty dr\, P(r,d, \eta)}, \end{equation} where $P(r,d,\eta)$ is the joint probability distribution in the catalog ({\it e.g.}, Strauss \& Willick 1995). The line-width is explicitly included to ensure that the correction holds when the selection criteria depend on $\eta$. This joint distribution is derived from several input quantities. One is the underlying spatial number density of galaxies, $n(r)$, which is taken from a self-consistent real-space reconstruction from the IRAS 1.2 Jy redshift survey (as described in Sigad {\it et al.\ } 1998). Another input is the distribution of galaxy diameters, $\Phi(D)$. One also needs as input the conditional probability $P(\eta,B,I|D)$, that a galaxy with a given $D$ will have a line-width $\eta$ and absolute magnitudes $B$ and $I$. We adopted the same distribution functions as those used in Freudling {\it et al.\ } (1995). Taking into account the selection in angular diameter, $a=D/r$, and the apparent blue magnitude limit $m_{B,{\rm max}}$, one obtains \begin{equation} P(r,d,\eta ) \propto r^2n(r)\int _{-\infty }^\infty da\ S(a|r)\ \Phi(a r)\ r\ \ {\rm exp} \left( -{[{\rm ln}(r/d)]^2 \over 2\Delta^2} \right)\ P(r,d,\eta,a r|m_{B,{\rm max}}) . \end{equation} The selection function of angular diameters at a given true distance, $S(a|r)$, is derived from the corresponding selection function in redshift space, $S_z(a|z)$, via \begin{equation} S(a|r) = \int _{-\infty }^\infty dz\ S_z(a|z) \ P(z|r), \end{equation} where $P(z|r)$ is based on the model peculiar-velocity field. The joint distribution $P(r,d,\eta,a r|m_{B,{\rm max}})$ is based on a combination of the diameter--magnitude relation, the correlation between B and I Magnitudes, the $\eta-B$ relation, and the B-magnitude limit in the selection of galaxies. This bias correction scheme was tested, and its details were refined, using carefully constructed mock catalogs (presented below, \S~\ref{subsec:mock}). We also tried several variants of the procedure to correct the real data for biases. The results of the power-spectrum analysis turn out to be fairly insensitive to the specifics of the bias correction scheme. In particular, for the underlying galaxy-density field that enters the correction via $n(r)$, we tried replacing the IRAS field (Sigad {\it et al.\ } 1998) with a linear reconstruction of a combination of IRAS and optical data (Freudling, da Costa \& Pellegrini 1994), and found negligible effects on the results of the likelihood analysis. The estimated errors in the observed TF relation can be directly translated into a distance uncertainty for each galaxy prior to the correction for biases. However, the correction for biases changes the properties of the scatter for a given location in estimated-distance space, which leads to a different uncertainty in the bias-corrected distance estimate. The semi-analytic approach is used also for a re-evaluation of the distance errors after the bias correction. We find that the bias correction acts towards slightly decreasing (by $\sim 10\%$) the average error, because of the additional information incorporated by the selection effects and the underlying density field used in the bias correction. The validity of this approach is verified using the mock catalogs (which also shows that the distribution of distance errors after the bias correction closely resembles a Gaussian distribution, Eldar {\it et al.\ } 1998). An independent verification of the magnitude of errors within the framework of the likelihood analysis is described in \S~\ref{subsec:tlcdmT60}. In what follows, we refer to these errors as our `original' error estimates. \section{METHOD} \label{sec:method} \subsection{Likelihood Analysis} \label{subsec:like} The goal of this paper is to estimate the power spectrum of mass density fluctuations from peculiar velocities, by finding maximum likelihood values for parameters of assumed model power-spectra. Again, the underlying assumptions are that the velocities and their errors are Gaussian, and that the velocity correlations can be derived from the density PS using linear theory. The assumption regarding the Gaussianity of the velocity field is supported by simulations which show that it is Gaussian well into the quasi-linear regime (Kofman {\it et al.\ } 1994). This is farther verified for our data set by the fact that the distribution of observed ${\rm ln}(z/d)$ closely resembles a Normal distribution. The validity of the second assumption is discussed later in \S~\ref{subsubsec:nlnl}. The likelihood analysis method is described in Zaroubi {\it et al.\ } (1997; see also Kaiser 1988; Jaffe \& Kaiser 1994). Here we summarize the main ideas, the underlying assumptions, and the specific application to peculiar velocities. Given a data set ${\bf d}$, our objective is to estimate the most likely model ${\bf m}$. Using Bayes theorem \begin{equation} \P ({\bf m} \vert {\bf d} ) = {\P({\bf m}) \P({\bf d}|{\bf m}) \over \P({\bf d})} \,, \end{equation} and assuming a uniform prior $\P({\bf m})$, this can be turned to maximizing the likelihood function, the probability of obtaining the data given the model, $\L=\P({\bf d}|{\bf m})$, as a function of the assumed model parameters. Under the assumption that both the underlying velocities and the observational errors are independent Gaussian random fields, the likelihood function can be written in the following form \begin{equation} \label{eq:like} {\cal L} = [ (2\pi)^N \det(R)]^{-1/2} \exp\left( -{1\over 2}\sum_{i,j}^N {u_i R_{ij}^{-1} u_j}\right)\,. \end{equation} This is simply the corresponding multivariate Gaussian distribution, where $\{u_i\}_{i=1}^{N}$ is the set of $N$ observed peculiar velocities at locations $\{{\bf r}_i\}$, and $R$ is their correlation matrix. Expressing each data point as the sum of the actual signal and the observational error $u_i=s_i+\epsilon_i$, the elements in the correlation matrix have two contributions \begin{equation} R_{ij}\equiv<u_i u_j>=<s_i s_j>+<\epsilon_i \epsilon_j>=S_{ij}+{\epsilon_i}^2 \delta_{ij}. \end{equation} The first term is the correlation of the signal, that is calculated from theory. The second term is the contribution of the distance errors, which are assumed to be uncorrelated. This should be true for the observational errors and the intrinsic scatter of the TF relation. We tested the impact of uncertainties in the bias correction, which might lead to correlated errors, by varying parameters of our bias model within the expected uncertainties. The changes in the results reported below are negligible compared to other systematic and random errors. For a given PS, the signal terms are calculated using their relation to the parallel and perpendicular velocity correlation functions, $\Psi_{\Vert}$ and $\Psi_{\perp}$, \begin{equation} S_{ij}=\Psi_{\perp}(r)\sin\theta_i \sin\theta_j + \Psi_{\Vert}(r)\cos\theta_i \cos\theta_j \, , \end{equation} where $r=\vert {\bf r} \vert=\vert {\bf r}_j-{\bf r}_i \vert$ and the angles are defined by $\theta_i=\hat{{\bf r}_i}\cdot\hat{{\bf r}}$ (G\'orski 1988; Groth, Juszkiewicz \& Ostriker 1989). In linear theory, each of these can be calculated from the PS, \begin{equation} \Psi_{\perp,\Vert}(r)= {H_0^2 f^2(\Omega)\over 2 \pi^2} \int_0^\infty P(k)\, K_{\perp,\Vert}(kr)\, dk \,, \end{equation} where $K_{\perp}(x) = j_1(x)/ x$ and $K_{\Vert}(x) = j_0-2{j_1(x)/ x}$, with $j_l(x)$ the spherical Bessel function of order $l$. The cosmological $\Omega$ dependence enters as usual in linear theory via $f(\Omega)\simeq \Omega^{0.6}$, and $H_0$ is the Hubble constant. The likelihood analysis is performed by assuming some parametric functional form for the PS. For each assumed PS, the correlation matrix $R$ is obtained and used to calculate the likelihood function (eq.~[\ref{eq:like}]). Exploring the chosen parameter space, we find the PS parameters for which the likelihood is maximized. (Note that since the model parameters appear also in the normalizing factor of the likelihood function, through $R$, maximizing the likelihood is {\it not} equivalent to minimizing the $\chi^2$.) The main computational effort is the calculation and inversion of the correlation matrix $R$ in each evaluation of the likelihood. It is an $N \times N$ matrix, where the number of data points $N$ is typically more than $1000$. Since the input data are peculiar velocities, the method essentially measures the combination $f(\Omega)^2P(k)$, and not directly the mass-density $P(k)$ by itself. This degeneracy between $\Omega$ and the PS can be broken when $\Omega$ enters explicitly into the functional form characterizing the PS shape, as in CDM models (\S~\ref{subsec:models}). Confidence levels are estimated by approximating $-2{\rm ln}\L$ as a $\chi^2$ distribution with respect to the model parameters. The likelihood analysis provides only relative likelihoods of different models. An absolute measure of goodness-of-fit can be provided, for example, by the value of the $\chi^2$ obtained with the parameter values associated with the maximum likelihood. A $\chi^2$ per degree of freedom of about unity would indicate that the model provides a good statistical description of the data. \subsection{Power Spectrum Models} \label{subsec:models} In order to perform the likelihood analysis, a specific parametric form for the PS is needed. For the main analysis of the paper, we use families of generalized CDM models normalized by the COBE 4-year data. The general form of these models is \begin{equation} \label{eq:cdm} P(k) = A_{COBE}(n,\Omega,\Lambda)\, T^2(\Omega,\Omega_B,h; k)\, k^n\,, \end{equation} where $A$ is the normalization factor and $T(k)$ is the CDM transfer function proposed by Sugiyama (1995, a slight modification of Bardeen {\it et al.\ } 1986): \begin{equation} \label{eq:Tcdm} T(k) = {{\rm ln}\left(1+2.34q) \right) \over 2.34q} \left[1+3.89q+(16.1q)^2+(5.46q)^3+(6.71q)^4\right]^{-1/4}\,, \end{equation} \begin{equation} q=k \left[ \Omega h \, \exp (-\Omega_b -(2h)^{1/2} \Omega_b/\Omega)\, (\, h\, {\rm Mpc^{-1}}) \right]^{-1}\,. \end{equation} These models include open universes with no cosmological constant, flat models with a cosmological constant ($\Omega+\Omega_\Lambda=1$), and tilted models with a large-scale power index $n$ that can be different from unity. The latter may include tensor fluctuations with tensor to scalar ratio of quadrupole moments of $T/S=7(1-n)$. The free parameters in the CDM models are $\Omega$, h and n. In all cases the baryonic density is set to be $\Omega_b=0.024 h^{-2}$ ({\it e.g.}, Tytler, Fan \& Burles 1996). For each model, the amplitude A is fixed by the COBE 4-year data. We followed the COBE normalization adopted in Zaroubi {\it et al.\ } (1997), who used the COBE DMR data (Hinshaw {\it et al.\ } 1996) to set the PS amplitude calculated by different authors (G\'orski {\it et al.\ } 1995; Sugiyama 1995; White \& Bunn 1995) for various cosmological CDM-like models. The calculation of Sugiyama (1995) was used as a reference. For models not studied by him the other results were used, after matching them to Sugiyama's in the commonly studied models. For a summary of the COBE normalization results see G\'orski {\it et al.\ } (1998). In addition, we use a different parameterization of the same power spectra, namely \begin{equation} \label{eq:gamma} P(k)= A\, k\, T^2(k), \quad T(k) = \Bigl( 1 + [ ak/\Gamma + (bk/\Gamma)^{3/2} + (ck/\Gamma)^2 ] ^{\nu} \Bigr)^{-1/\nu}\,, \end{equation} with $a=6.4\,{\rm h^{-1}Mpc}$, $b=3.0\,{\rm h^{-1}Mpc}$, $c=1.7\,{\rm h^{-1}Mpc}$ and $\nu = 1.13$ ({\it e.g.}, Efstathiou, Bond \& White 1992). In the context of the CDM model, $\Gamma$ has a specific cosmological interpretation, $\Gamma=\Omega h$. Below, however, we use equation (\ref{eq:gamma}) as a generic form with limiting logarithmic slopes $n=1$ and $-3$ on large and small scales respectively, and with a turnover at some intermediate wavenumber that is determined by the single shape parameter $\Gamma$. Hereafter, we refer to this functional form of the power-spectrum as the ``$\Gamma$ model''. We use it as a convenient parameterization, for comparability with other works, and for relaxing the COBE normalization. The free parameters that we vary in this case are the amplitude $A$ and the shape parameter $\Gamma$. \subsection{Error Models} \label{subsec:errors} We make a special effort to estimate the distance errors. As mentioned in section~\ref{sec:intro}, this is done because the amplitude of the deduced PS depends on their sum in quadrature; if errors are overestimated, the PS is underestimated and vice versa. We first apply the likelihood analysis with the original distance errors, $\sigma_{oi}$, as estimated a priori for each galaxy ($i$) in the SFI catalog with the procedure explained in section~\ref{subsec:bias}. Alternatively, we incorporate the errors in the likelihood analysis itself, by allowing a parametric model for the errors in addition to the parametric model of the PS. An error model is fully specified by the standard deviations $\sigma_i$ because we assume that the distance errors for the individual galaxies are uncorrelated and that the scatter is Gaussian. We try two alternative global modifications of the original errors as our error model: one is based on a free multiplicative factor, $\sigma_i = p \sigma_{oi}$, and the other is based on a free additive constant in quadrature, $\sigma_i = (\sigma_{oi}^2 \pm q^2)^{1/2}$. The latter is similar to the way we modeled the difference in scatter between the field and cluster samples (see \S~\ref{subsec:derrors}). The errors are incorporated in the model that constitutes the correlation matrix, and the parameters $p$ or $q$ are adjusted simultaneously with the parameters of the PS until the likelihood is maximized. The apparent cost of adding the error parameter to the likelihood analysis is a larger formal error in the final results for the power spectrum and the cosmological parameters. However, since our original error estimates carries some uncertainty, this procedure, which provides an almost independent estimate of the errors, could add to the overall confidence in our results. \section{TESTING THE METHOD} \label{sec:mock} \subsection{Mock Catalogs} \label{subsec:mock} It is essential to check the method with realistic mock catalogs, in view of the large errors in the data and the approximations made in the analysis. For this purpose, we use the $N$-body simulation of Kolatt {\it et al.\ } (1996) which was designed to mimic the large-scale density distribution in our local universe. The simulation is based on initial conditions extracted from a reconstruction of the smoothed ($5\,{\rm h^{-1}Mpc}$ Gaussian) real-space density field from the IRAS 1.2 Jy redshift survey, taken back into the linear regime. Small-scale perturbations were added by means of constrained random realizations, and the system was then evolved forward in time using a particle-mesh $N$-body code until a present epoch defined by $\sigma_8=0.7$. The ``true'' PS was calculated directly from the underlying mass distribution of the simulation, by Fourier transforming to k-space and calculating the power in bins of wavelength. ``Galaxies" were identified in the simulation via a linear biasing scheme, and then divided into galaxy types, S's and E's, while obeying the morphology-density relation. Observational parameters were assigned to the S galaxies in the mock sample according to the prescription of Freudling {\it et al.\ } (1995), and perturbed at random according to the estimated observational errors. Subsequently, we selected ten random mock SFI samples using the exact selection criteria of the real SFI sample. Each of these mock catalogs was corrected for biases, and the errors were re-evaluated accordingly, in the same way as in the real data (see Eldar {\it et al.\ } 1998). \subsection{Testing with the $\Gamma$ Model} \label{subsec:mock_gamma} We first apply the likelihood analysis to the mock SFI catalogs using the $\Gamma$ functional form (eq.~[\ref{eq:gamma}]) as the prior model for the PS. We allow the amplitude $A$ and the shape parameter $\Gamma$ to vary, and include an additional free parameter in the error model. It is realized that the freedom provided by this family of models (just as by any other family of models) may not be enough for an adequate fit to the true PS. No additional constraint is applied on large scales, so this is a test of the ability of the velocity data alone to constrain the PS. Figure~\ref{fig:mockg} (left panel) shows a contour plot of the resulting log-likelihood (${\rm ln} \L$) in the parameter plane ($A-\Gamma$), as obtained from one of the realizations of the mock catalogs. The errors in this case were allowed to vary by the multiplicative factor $p$, and the plot shown corresponds to the best-fit error parameter. Here, and in all the figures that follow, the log-likelihood contours are relative to the maximum likelihood with contour spacing of $\Delta[{\rm ln} \L]=-1$. The right panel of Figure~\ref{fig:mockg} shows the corresponding best-fit power spectrum (solid line). The filled symbols mark the target of the reconstruction --- the true PS of the simulation. The shaded area about the derived PS corresponds to the region of $90\%$ confidence about the most-likely parameters in the likelihood plot, for fixed errors. The uncertainty becomes large at small $k$'s corresponding to scales larger than the sampled volume, because no additional data were used to constrain the PS on large scales. The figure demonstrates that for this random realization the likelihood analysis with the $\Gamma$ model recovers the true PS well within the error-bars. A similar quality of recovery is obtained for all the random realizations of the mock catalogs, and also when the errors are varied in the alternative way. \begin{figure}[tbp] {\special{psfile="f1.ps" angle= 0 hscale=90 vscale=90 hoffset=-38 voffset=-605}} \vspace{7.7 cm} \caption{\capt Left panel: Contour map of ${\rm ln} \L$ in the $A-\Gamma$ plane for one random mock catalog. Contour spacing is $\Delta[{\rm ln} \L]=-1$. $A$ is in units of $A_0 = 2.0 \times 10^6 (\,{\rm h^{-1}Mpc})^4$. The errors were varied by a multiplicative factor. Right panel: The PS corresponding to the maximum-likelihood $\Gamma$-model parameters determined for the mock catalog. The filled symbols mark the true PS of the simulation. The shaded area around the PS is the $90\%$ confidence region for the best-fit errors, obtained from the contour map. } \label{fig:mockg} \end{figure} The maximum-likelihood errors are found to be within $5\%$ of their ``true" values. The latter were estimated by slightly modifying the known distance errors (as built into the mock catalogs) after correcting for Malmquist bias. The $5\%$ error reflects the imperfect match between the assumed family of shapes for the PS and the true shape, and, perhaps, the uncertainty in the modification of the error estimate or the slight deviation of the modified errors from a Gaussian distribution. \subsection{Testing with a Tilted \lcdm Model} \label{subsec:mock_tlcdm} We wish to check the success of the likelihood analysis also with the COBE-normalized CDM models. We choose as our test case the flat ($\Omega+\Omega_\Lambda=1$) \lcdm family of models, with tensor fluctuations, a corresponding tilt in $n$, and a Hubble constant of $h=0.6$. COBE normalization is imposed as if the mock simulation is identical to the real universe. The likelihood analysis is thus performed by varying the parameters $\Omega$, $n$ and the error-parameter $p$ or $q$. This family of shapes for the PS is, again, not guaranteed to provide a perfect fit to the true PS. In particular, the parameter-dependent COBE normalization is not guaranteed to give the correct amplitude, since the simulation was not explicitly constrained to produce the level of large-scale CMB anisotropies detected in the real universe. Figure~\ref{fig:mockt} (right panel) shows the best-fit power spectra of the 10 mock SFI catalogs, superimposed on the true PS. This test uses the $q$ error parameter. The left panel shows ${\rm ln} \L$ contours in the $\Omega-n$ plane for one representative mock catalog, with the maximum-likelihood points for all ten catalogs marked. We see that all the maximum-likelihood points fall along the ridge of high-likelihood in the one case plotted, and are therefore moderately consistent with one another. A way to translate the likelihood contours to errors in the values of the model parameters is by assuming that, with the errors fixed, $-2{\rm ln}\L$ has a $\chi^2$ distribution with two degrees of freedom. Then, the $1\sigma$ confidence level around the maximum-likelihood point is at ${\rm ln}\L \sim -1.15$ and the $90\%$ confidence level is at ${\rm ln}\L \sim -2.3$. The fact that indeed six of the ten cases fall within the $1\sigma$ contour as determined above, and nine cases fall within the $90\%$ confidence level, indicate that this crude error estimate is quite reasonable. The $90\%$ confidence region for this specific catalog is again drawn as a shaded area in the PS plot; one can see that this region indeed resembles the actual scatter of the ten cases. The maximum-likelihood power-spectra fit reasonably well the true PS, with a fairly small spread on small and intermediate scales. For large scales (small $k$'s) the scatter is somewhat larger, but not as large as for the $\Gamma$ model which was completely free at large scales. Again, the success of recovery is similar when the alternative error parameter is used, and the errors are similar to those obtained in the case of the $\Gamma$ model. It is encouraging to note that the recovery of the PS is fairly robust on the relevant scales ($k\sim 0.1\, h\, {\rm Mpc^{-1}}$) among the realizations, and independent of the prior model assumed for the PS, or the assumed error model. \begin{figure}[tbp] {\special{psfile="f2.ps" angle= 0 hscale=90 vscale=90 hoffset=-38 voffset=-605}} \vspace{7.7 cm} \caption{\capt Left panel: Contour plot of ${\rm ln}\L$ in the $n-\Omega$ plane for one of the mock catalogs, for the tilted \lcdm model with tensor fluctuations and $h=0.6$, and the errors varied in quadrature. The best-fit values for all catalogs are marked by `+'. Right panel: Best-fit PS of the 10 mock catalogs (thick lines representing same curves derived for two different catalogs). The shaded area represents the $90\%$ confidence region for the catalog whose contour plot is shown. The filled symbols mark the true PS of the simulation. } \label{fig:mockt} \end{figure} \section{RESULTS} \label{sec:results} \subsection{Maximum-Likelihood Errors} \label{subsec:tlcdmT60} Before estimating the PS from the actual SFI data, we investigate the reliability of our observational error estimate by allowing certain freedom in the errors. As a test case we use as a prior for the PS the COBE-normalized \lcdm family of models, with tensor fluctuations and a corresponding tilt in $n$, and with the Hubble constant fixed at $h=0.6$. We perform the likelihood analysis on the real SFI data varying $\Omega$ and $n$, with the errors treated in three different ways; first with the errors fixed at their original values, and then by varying them according to the two error models discussed above. Figure~\ref{fig:tlcdmT60} summarizes the results obtained in these cases. The left panel shows the ${\rm ln}-$likelihood contours in the $\Omega-n$ plane for the case of fixed errors, with the best-fit points for the three cases marked. The corresponding power spectra are presented in the right panel. In the $p$ case, the preferred errors are $5\%$ larger than the original ones, while in the $q$ case the preferred errors are smaller by 0.03 in quadrature (typically a decrease of $\sim 2\%$). \begin{figure}[t] {\special{psfile="f3.ps" angle= 0 hscale=90 vscale=90 hoffset=-38 voffset=-605}} \vspace{7.6 cm} \caption{\capt Left panel: Contour plot of ${\rm ln}\L$ in the $\Omega-n$ plane for the SFI sample, for the tilted \lcdm model with tensor fluctuations and $h=0.6$, using the original error estimates. The best-fit point is marked with a `+'. The maximum likelihood locations when varying the errors in quadrature (`$q$') or by a multiplicative factor (`$p$') are also marked. Right panel: The most-likely PS for this model, for these three variants of the errors. The dotted region around the PS represents the $90\%$ confidence limit for the case of the original errors (thick line), obtained from the high-likelihood ridge shown in the contour map. } \label{fig:tlcdmT60} \end{figure} The different trends in the likely errors reflect our uncertainty of the exact form of the error model. We note that while these changes are in opposite directions, they are of small magnitude, within the uncertainty expected based on the mock catalogs. The corresponding changes in the best-fit parameters are along the ridge of high likelihood in the $\Omega-n$ plane, within the $1\sigma$ confidence level, {\it i.e.}, it is hardly significant. In all three cases, $\chi^2/N_{dof} \sim 1$ for the best-fit PS ($1.02$ for the original, fixed errors, $0.99$ for the p error model and 1.02 for the q error model), implying that all are reasonable fits to the data. Similar results concerning the errors are obtained when the other PS models are used as priors. The error estimate is robust to variations in the original errors about which the error model is perturbed. This likelihood analysis of the errors thus provides a very encouraging indication that the original error estimates in SFI are accurate to better than $5\%$. Note that ``original'' here refers to the refined SFI errors after the correction for biases (\S~\ref{subsec:bias}). The fact that the likelihood analysis and the semi-analytic correction converge to the same error estimate is encouraging. Based on this finding, we perform the rest of the analysis in this paper using fixed errors at their original values. \subsection{COBE Normalized CDM Models} \label{subsec:cdm} We now use the generalized CDM families of cosmological models of the form described in equations (\ref{eq:cdm}) and (\ref{eq:Tcdm}). Our models include open CDM (\ocdm), flat models with a cosmological constant, and tilted models with or without tensor fluctuations, allowing for variations in the cosmological parameters $\Omega$, $h$, and $n$. For each specific choice of model and parameters the amplitude is fixed according to the 4-year COBE normalization. \subsubsection{Scale-Invariant Models} \label{subsubsec:si_cdm} \begin{figure}[tbp] {\special{psfile="f5.ps" angle= 0 hscale=90 vscale=90 hoffset=-35 voffset=-605}} \vspace{7.6 cm} \caption{\capt Contour plot of ${\rm ln}\L$ in the $\Omega-h$ plane for scale-invariant \ocdm model (left panel) and \lcdm model (right panel). The most-likely value of $\Omega$ and its $90\%$ error bar are marked for a fixed value of $h=0.6$. } \label{fig:olcdm} \end{figure} Figure~\ref{fig:olcdm} shows the resulting likelihood contours for the scale-invariant case, $n=1$, for the \ocdm model and the \lcdm model. The contours are plotted in the $\Omega-h$ plane. The best-fit parameters in each case are marked, but it is clear from the elongated contours that the two parameters are not determined separately. The high-likelihood ridges rather constrain a degenerate combination of these parameters, which can be roughly fitted by the following functions: \begin{eqnarray} \label{eq:ocdm} \Omega\, {h_{60}}^{0.9} & = & 0.68\pm 0.06\,, \qquad \rocdm; \\ \label{eq:lcdm} \Omega\, {h_{60}}^{1.3} & = & 0.59\pm 0.07\,, \qquad \rlcdm \,. \end{eqnarray} The error-bars (here, and throughout the paper) arise from the joint $90\%$ confidence region of the parameters. The constraints on $\sigma_8 f(\Omega)$, obtained by integrating over the corresponding power spectra, are ${0.83}^{+0.07}_{-0.11}$ and ${0.81}^{+0.13}_{-0.07}$ for \ocdm and \lcdm respectively. The error-bars quoted are the marginalized 1-dimensional $90\%$ confidence limits. For an assumed value of $h$, {\it e.g.}, $h=0.6$, the maximum-likelihood values are $\Omega=0.67 \pm 0.05$ for \ocdm and $\Omega=0.58 \pm 0.06$ for \lcdm (as marked on the plots). The $\chi^2/N_{\rm dof}$ for the best-fit PS are $1.01$ and $1.04$ respectively, with similar values along the high-likelihood ridge. With $N_{\rm dof}=1213$, one expects for a good fit $1.00\pm0.04$, so our CDM models indeed provide good fits to the data. \subsubsection{Tilted Models} \label{subsubsec:tcdm} Figure~\ref{fig:tolcdm} presents the results obtained when allowing for a tilt in the PS on large scale relative to $n=1$, both for the \ocdm and \lcdm families of models. The first cases considered are with scalar fluctuations only, $T/S=0$. We fix the Hubble constant here at $h=0.6$ while varying $\Omega$ and $n$. Again, the elongated ridge of high-likelihood determines a certain degenerate combination of the parameters, which can be approximated by: \begin{eqnarray} \label{eq:tocdm} \quad \Omega \, n^{1.4}\,({h_{60}}^{0.9})& = &0.68\pm 0.07\, ,\qquad\rocdm ;\\ \label{eq:tlcdm} \quad \Omega \, n^{2.0}\,({h_{60}}^{1.3} )& = &0.58\pm 0.08\, ,\qquad\rlcdm \, . \end{eqnarray} The $h$ dependence is determined for the $n=1$ case. The corresponding constraints are $\sigma_8 f(\Omega)={0.83}^{+0.08}_{-0.10}$ for the tilted \ocdm case and $\sigma_8 f(\Omega)={0.82}^{+0.10}_{-0.09}$ for tilted \lcdm. The $\chi^2/N_{\rm dof}$ values are $1.02$ in both cases, again a good fit. \begin{figure}[tbp] {\special{psfile="f6.ps" angle= 0 hscale=90 vscale=90 hoffset=-35 voffset=-605}} \vspace{7.6 cm} \caption{\capt Contour plot of ${\rm ln}\L$ in the $\Omega-n$ plane, for the tilted \ocdm model (left panel) and tilted \lcdm model (right panel). In both cases $h=0.6$ and no tensor component is included. } \label{fig:tolcdm} \end{figure} The case of the tilted \lcdm family of models, with $h=0.6$ and with a tensor component of $T/S=7(1-n)$, was partly discussed already as our default case in \S~\ref{subsec:tlcdmT60}. The likelihood map in Figure~\ref{fig:tlcdmT60} reveals the familiar situation of a high-likelihood ridge that constrains a degenerate combination of the cosmological parameters, now approximated by \begin{eqnarray} \label{eq:tlcdmT60} \qquad \Omega\, n^{3.9} ({h_{60}}^{1.3})& = &0.58\pm 0.08\,, \qquad \rlcdm\,+{\rm tensor}\,. \end{eqnarray} The $h$ dependence is determined for $n=1$. The corresponding value of $\sigma_8 f(\Omega)$ is $0.81^{+0.09}_{-0.08}$. The uncertainty associated with the PS, shown as the shaded area in the right panel of Fig.~\ref{fig:tlcdmT60}, is similar to the uncertainty in the other COBE-normalized CDM variants. \subsection{$\Gamma$ Model} \label{subsec:gamma} Finally, we use the $\Gamma$ model as a prior for the PS, varying the amplitude $A$ and shape parameter $\Gamma$ with no additional constraints imposed at large scales. Figure~\ref{fig:gamma} shows the contours of ${\rm ln} \L$ in the $A-\Gamma$ plane, and the corresponding best-fit PS. The maximum likelihood values are $\Gamma=0.375 \pm 0.14$ and $A = 5.0\times 10^5 (\,{\rm h^{-1}Mpc})^4$. The $\chi^2$ per degree of freedom for the maximum likelihood parameters is $\chi^2/N_{\rm dof}=1.03$, indicating that this is a good fit to the data. The constraint obtained by integrating over the power spectra is $\sigma_8 f(\Omega) = {0.80}^{+0.09}_{-0.08}$. The scatter at small $k$'s is larger than in the case of the COBE-normalized models, due to the amplitude freedom, as seen already in the mock catalogs. \begin{figure}[t] {\special{psfile="f4.ps" angle= 0 hscale=90 vscale=90 hoffset=-38 voffset=-605}} \vspace{7.6 cm} \caption{\capt Left panel: Contour plot of ${\rm ln}\L$ for the $\Gamma$ model. The best-fit point is marked with a `+'. Right panel: The best-fit PS, with the shaded area marking the uncertainty based on the $90\%$ confidence region of the likelihood contours. } \label{fig:gamma} \end{figure} \section{ROBUSTNESS OF RESULTS} \label{sec:robust} The error estimates in the parameters given in the previous section are formal $90\%$ confidence levels. In this section, we test the robustness of these results to various variations in the data and models used. \subsection{Robustness to Models} \label{subsec:robust} Figure~\ref{fig:all} shows the power spectra corresponding to the maximum-likelihood parameters for all the models presented so far in this paper, including the COBE-normalized CDM variants and the $\Gamma$ model. The $90\%$ confidence region for the tilted \lcdm model with tensor fluctuation and $h=0.6$ is drawn as well, as a reference for the uncertainty associated with each model based on the likelihood contours. The similarity of all the curves is striking; they agree well within the formal uncertainties of each other. The agreement is excellent for $k>0.1$, on the scale where the data constrain the models effectively. The difference between the curves shows a slightly larger scatter on larger scales, not properly sampled by the present data. The similarity of the results using as priors the COBE-normalized CDM models and the amplitude-free $\Gamma$ model indicates that the peculiar velocity data themselves contain meaningful information to constrain the PS. \begin{figure}[tbp] {\special{psfile="f7.ps" angle= 0 hscale=51 vscale=51 hoffset=65 voffset=-312}} \vspace{7.6 cm} \caption{\capt The maximum-likelihood power spectra based on the $\Gamma$ model (solid line) and the various COBE-normalized CDM models. The shaded area is once again the $90\%$ confidence region for the tilted \lcdm model with tensor fluctuations and $h=0.6$. } \label{fig:all} \end{figure} Table~1 summarizes the features of the most likely power spectra based on the various prior models. The approximate constraint on the combination of cosmological parameters as obtained from the high-likelihood ridge is given for each case. The best-fit values of the individual cosmological parameters are also listed, but recall that they carry large uncertainties. The exact location of the maximum-likelihood point in the high-likelihood ridge is hardly significant. Parameters that were held fixed in the likelihood analysis are marked in brackets. Several characteristics of the best-fit power spectra are listed: the value of $\sigma_8 f(\Omega)$, the amplitude of $f(\Omega)^2 P(k)$ at $k=0.1 \,{\rm h^{-1}Mpc}$ and the location of the PS peak, $k_{peak}$. The error-bars quoted in the header represent the typical $90\%$ confidence uncertainty in these quantities within each family of models. \begin{table}[tb] \label{table} \caption{Maximum-Likelihood Results for the various models} \vspace{0.5 cm} \begin{tabular}{@{}c@{ }c@{}c@{}c@{}c@{}c@{ }@{ }c@{ }@{ }c@{}c@{ }c@{}} \multicolumn{10}{c}{COBE-normalized CDM models}\\ \hline \hline \vspace{0.2 cm} & & $\sigma_8\Omega^{0.6}$ & $P_{0.1}\Omega^{1.2}$ & $k_{peak}$ & & & & & \\ & High Likelihood & & \small{($h^{-3}Mpc^3$)} & \small{($hMpc^{-1}$)} & & & & & \\ CDM Model & Ridge & \small{($\pm0.10$)} & \small{($\pm1500$)} & \small{($\pm0.01$)} & $\Omega$ & $n$ & $h$ & $\chi^2/N_{\rm dof}$ & $-{\rm ln}{\cal L}$ \\ \hline \vspace{0.4 cm} Open, n=1 & $\Omega {h_{60}}^{0.9}=0.68\pm 0.06$ & 0.83 & 4400 & 0.038 & 0.55 & (1) & 0.78 & 1.01 & 8579.4 \\ \vspace{0.4 cm} $\Lambda$, n=1 & $\Omega{h_{60}}^{1.3}=0.59\pm 0.07$ & 0.81 & 4600 & 0.031 & 0.95 & (1) & 0.42 & 1.04 & 8580.1 \\ \vspace{0.4 cm} Tilted-Open & $\Omega n^{1.4}=0.68\pm 0.07$ & 0.83 & 4600 & 0.035 & 0.72 & 0.96 & (0.6) & 1.02 & 8579.5 \\ \vspace{0.4 cm} Tilted-$\Lambda$ & $\Omega n^{2.0}=0.58\pm 0.08$ & 0.82 & 4200 & 0.037 & 1.00 & 0.76 & (0.6) & 1.02 & 8579.6 \\ Tilted-$\Lambda$ & $\Omega n^{3.9}=0.58\pm 0.08$ & 0.81 & 4300 & 0.037 & 0.79 & 0.92 & (0.6) & 1.02 & 8579.5 \\ +tensor & & & & & & & & & \\ \hline $\Gamma$ model & $\Gamma=0.375\pm 0.14$ & 0.80 & 4300 & 0.037 & & & & 1.03 & 8579.3 \\ \end{tabular} \end{table} The typical results for the PS are $P(k=0.1 \, h\, {\rm Mpc^{-1}}) \Omega^{1.2} =(4.4\pm1.5)\times 10^3 \, ( h^{-1} {\rm Mpc})^3$ and $\sigma_8 \Omega^{0.6} = 0.82 \pm 0.10$. The variations from model to model are much smaller than the formal errors for each model, increasing the above errors to $1.7$ and $0.12$ respectively. These results are thus almost independent of the model, at least for the family of models considered here. The actual likelihood values of all the best-fit models are very similar, and all have comparable $\chi^2/N_{dof} \simeq 1$ values. The variation of the high-likelihood ridge between the \lcdm and \ocdm families of models is more noticeable. The general constraint on the combination of cosmological parameters can be roughly approximated by $\Omega\, n^{\nu}\, {h_{60}}^{\mu} = 0.62 \pm 0.15$, where the error includes the formal uncertainties of the three parameters and the variations between models. For \lcdm, $\mu=1.3$ and $\nu=2.0,3.9$ without and with tensor fluctuation respectively. For \ocdm, without tensor fluctuations, the powers are $\mu=0.9$ and $\nu=1.4$. The similarity of the power spectra obtained using the COBE-normalized CDM models and the COBE-free $\Gamma$ model (see Table~1) indicates that the PS is predominantly determined by the velocity data. Therefore, we have so far ignored the error associated with the COBE normalization. As a test for the sensitivity to this error, we have repeated the analysis using the tilted \lcdm model (with tensor fluctuations), but now normalized alternatively $\sim 18\%$ higher or lower then the mean COBE values (in accordance with the relative $\pm 1 \sigma$ uncertainty associated with $Q_{rms-PS \vert n=1}$, Bennett {\it et al.\ } 1996). This results in a slight shift of the high-likelihood ridge, corresponding to a $\sim 6\%$ change in the constraint on the combination of parameters (eq.~[\ref{eq:tlcdmT60}]; a smaller value is obtained for the higher normalization and vice versa), which is within our formal $1 \sigma$ error-bars. However, the combined effect of the different amplitude and corresponding cosmological parameters on the PS is essentially negligible, with $\sigma_8 \Omega^{0.6}$ varying by only $0.01$. \subsection{Zero-point Uncertainty} \label{subsec:0p} \begin{figure}[t] {\special{psfile="f8.ps" angle= 0 hscale=90 vscale=90 hoffset=-38 voffset=-650}} \vspace{7.7 cm} \caption{\capt Left panel: Contour plots in the $\Omega-n$ plane for the tilted \lcdm $h=0.6$ model with a tensor component, for the original calibration. The maximum-likelihood point is marked by `+'. The maximum-likelihood values when varying the global zero-point by a $\pm 5\%$ Hubble flow are marked by `+5',`-5' accordingly. Right panel: The most likely PS for the original zero-point calibration (solid line) and when varying the zero-point by $\epsilon=+0.05$ (long dashed line) and $-0.05$ (short dashed line). The corresponding $\sigma_8 \Omega^{0.6}$ values are marked on the plot. The shaded region is the formal likelihood $90\%$ confidence region for the original case. } \label{fig:0point} \end{figure} A fundamental freedom in the measured peculiar velocities is in the global zero-point of the TF relation, which fixes the distances at absolute values (in $\ {\rm km\,s^{-1}}$). Changing the zero-point, that is multiplying the distances $r$ by a factor $(1-\epsilon)$, is equivalent to adding a monopole Hubble-like flow $\epsilon r$ to the peculiar velocities. The zero-point calibration of the TF relation used for the SFI sample was obtained from the SCI catalog of $\sim 500$ galaxies within $24$ clusters, using the ``Basket of Clusters'' approach (Giovanelli {\it et al.\ } 1997a, 1997b). The uncertainty in the zero point was estimated to be about 0.05~magnitudes, which corresponds to an uncertainty in the velocity field of $2.5\%$ of the distance. To estimate the effects of such uncertainties we have run the likelihood analysis with our tilted \lcdm test-case, conservatively using zero-point changes of twice the estimated uncertainty, $\epsilon=\pm 0.05$. Figure~\ref{fig:0point} illustrates the effect on the results for these cases. The changes of zero-point appear to shift the location of the maximum-likelihood values essentially along the high-likelihood ridge. The high ridge is not altered by much when the zero point varies in this range. The right panel shows the resulting best-fit PS for the three different zero points, and lists the corresponding values of $\sigma_8 \Omega^{0.6}$. While the variations in the zero-point systematically affect the PS, the changes are not large; they fall within the range of the formal likelihood errors, and are of the same order of the uncertainty associated with the random distance errors (compare to Fig.~\ref{fig:tlcdmT60}). It is encouraging that the amplitude of the PS on intermediate scales ($k \sim 0.1 \, h\, {\rm Mpc^{-1}}$) is robust vis-a-vis changes in in the zero point. Similar results were obtained when using the other families of PS prior models. Similar to the uncertainty in the zero point of the distance indicator, there is also an uncertainy associated with the slope of the TF relation. This could lead to correlated errors in the inferred distances and peculiar velocities, due to the fact that the average linewidth of SFI galaxies slightly depends on distance (Wegner {\it et al.\ } 1999). However, the impact of this uncertainty on our results is even slightly smaller than that of the uncertainty in the zero point, perhaps because the SFI sample has been selected intentionally to minimize the distance dependence of the line widths. \subsection{Nonlinear Effects} \label{subsec:nl} A basic assumption in our analysis has been that linear gravitational instability theory is adequate for the purpose of recovering the PS from observed velocities on the scales of interest here. This is based on the fact that in the mildly-nonlinear regime the velocity field is approximated by linear theory better than the density field (basically because the velocity is a spatial integral of the density and is affected by fluctuations on larger scales). Indeed, the success of the recovery of the PS from the mock catalogs leads us to believe that this assumption is justified. However, one cannot rule out the possibility that some nonlinear effects are artificially reduced to some degree in the particle-mesh $N$-body simulation, and it is possible that the smooth shape of the linear PS as predicted for the CDM family of models may fail to properly match the nonlinear features that may be present on small scales in the real data. Therefore, we discuss in this section possible nonlinear effects, which could manifest themselves in different forms. For example, as coherent motions associated with the non-linear evolution of the PS (\S~\ref{subsubsec:nlnl}), or as incoherent random motions, perhaps due to shell-crossing, which may be modeled as an additional velocity component of dispersion $\sigma_v$ (\S~\ref{subsubsec:sigmav}). \subsubsection{Nonlinear Power Spectra} \label{subsubsec:nlnl} A way to include more properly nonlinear effects in our analysis is by developing an approximation for the nonlinear evolution of the PS and then incorporating it in the likelihood analysis. Such approximations exist for the density power spectrum, $P_\delta$ ({\it e.g.}, Peacock \& Dodds 1994; Jain, Mo \& White 1995; Peacock \& Dodds 1996, hereafter PD), but we need a similar approximation for the evolution of the {\it velocity} power spectrum, $P_v$, which is the quantity we actually confront with the data. A development and application of such an approximation is beyond the scope of the present paper and will be presented later (Zehavi {\it et al.\ } 1999). Here, we summarize some relevant issues and illustrate the magnitude of such effects. Figure~\ref{fig:pvk} shows the velocity PS computed in several different ways from an adaptive P$^3$M cosmological $N$-body simulation with a resolution higher by an order of magnitude than the simulation used in the present paper for the mock catalogs (but inside a smaller box of size $85 \,{\rm h^{-1}Mpc}$; GIF simulation, Colberg {\it et al.\ } 1999). The initial model used for $P_\delta$ is the so-called $\tau$CDM model with $\Omega=1.0$, $h=0.5$ and a modified shape parameter $\Gamma=0.21$. The figure clearly demonstrates that the velocity PS is reproduced by linear theory much better than the density PS. The $P_v$ that is computed directly from the evolved velocity field of the simulation (solid dots) lies slightly below the $P_v$ obtained from the assumed $P_\delta$ using linear theory ($P_v \propto k^{-2} P_\delta$, solid line). On the other hand, the nonlinear correction to $P_\delta$ ({\it e.g.}, PD) is larger than that of $P_v$ and in the opposite direction (upwards, as can be seen by the open dots and dashed line in Fig.~\ref{fig:pvk}). One might have naively expected that the likelihood analysis using a pure linear treatment would be inferior to incorporating a non-linear correction for $P_\delta$ followed by a linear translation to $P_v$. However, as illustrated in Figure~\ref{fig:pvk}, this is not the case. The latter procedure overestimates the nonlinear effects on the velocity PS and increases the bias in the results. A similar bias is reproduced when using the mock catalogs from the low-resolution simulation of Kolatt {\it et al.\ } (1996), which exhibits a similar behavior as in Fig.~\ref{fig:pvk}. This could be remedied, in principle, by incorporating the evolved $P_\delta$ and then counter-balancing it with a proper approximation for the nonlinear velocity--density relation, but this would be risky as we would be applying two large corrections in opposite directions to mimic a small net effect. Until we develop a direct nonlinear correction for $P_v$, we adopt the fully linear procedure as our best approximation. This is justified by its success in the mock catalogs and by the expectation for only small nonlinear effects in $P_v$. \begin{figure}[tbp] {\special{psfile="f10.ps" angle= 0 hscale=51 vscale=51 hoffset=65 voffset=-312}} \vspace{7.6 cm} \caption{\capt The velocity PS as computed from the high-resolution simulation of Colberg {\it et al.\ } (1999) (solid dots), compared with the theoretical linear PS (solid line), the corrected PS using the PD formalism (dashed line) and a computation via the density PS of the simulation (open dots). The latter three have been transformed to velocity PS using the linear velocity--density relation. } \label{fig:pvk} \end{figure} \subsubsection{Random Motions} \label{subsubsec:sigmav} We have made an ad-hoc attempt to model non-linearities by introducing an uncorrelated velocity component of constant dispersion $\sigma_v$, that adds a free term at zero-lag to the correlation function derived from the linear PS model. This may be a crude way to represent small-scale random motions that are associated with multi-streaming. An alternative interpretation of this additional parameter may be as an unrecognized uncertainty in the distance estimate which does not depend on distance and is therefore not included in our usual error model. In either case, this provides a test for the robustness of our results to an additional degree of freedom. Figure~\ref{fig:sigma_v} demonstrates the effect of including a free $\sigma_v$ in the likelihood analysis, again for our tilted \lcdm test-case. When allowing for this extra freedom, the preferred value turns out to be $\sigma_v=200 \pm 120 \ {\rm km\,s^{-1}}$, and is associated with a PS that is slightly lower for $k>0.1$ and somewhat higher at small $k$. The value of $\sigma_8 \Omega^{0.6}$ is reduced by $14\%$. The deviations, in general, are comparable to the formal likelihood $90\%$ uncertainty marked by the shaded area. The likelihood contours are somewhat sparser in this case, because of the additional scatter that reduces the sensitivity to variations in the parameters. The ridge of high likelihood is slightly shifted toward smaller values of the cosmological parameters and it can now be roughly described by \begin{eqnarray} \label{eq:tlcdmT60_sigv} \qquad \Omega\, n^{3.9} ({h_{60}}^{1.3})& = &0.49\pm 0.09\,, \end{eqnarray} a $\sim 15\%$ decrease compared to equation (\ref{eq:tlcdmT60}), which is of the order of the random error. \begin{figure}[tbp] {\special{psfile="f9.ps" angle= 0 hscale=90 vscale=90 hoffset=-38 voffset=-605}} \vspace{7.7 cm} \caption{\capt Left panel: Likelihood contour plot in the $\Omega-n$ plane for the tilted \lcdm $h=0.6$ model with tensor fluctuations, with an additional scatter of $\sigma_v=200 \ {\rm km\,s^{-1}}$. The maximum-likelihood point is marked by `$*$', and the corresponding point for $\sigma_v=0$ is marked by `+'. Right panel: The most likely PS when including $\sigma_v=200$ in the fit (dashed line) and for the original $\sigma_v=0$ case (solid line) together with its $90\%$ confidence region shaded. } \label{fig:sigma_v} \end{figure} Such a preference for a non-zero $\sigma_v$ associated with a change in the PS is not recovered in the mock SFI catalogs, for which a similar likelihood analysis turns out to prefer a negligible $\sigma_v$ and a negligible effect on the PS. $N$-body simulations of higher resolution may clarify this situation. We note that the inclusion of a free $\sigma_v$ in the fit to the real SFI data leads to results similar to those obtained when including a free multiplicative parameter in the error model (\S~\ref{subsec:tlcdmT60}). The interpretation of a nonzero $\sigma_v$ is thus not unique: It may refer to nonlinear effects that exist in the real data but not in the current simulation, or it may indicate that the actual errors are slightly larger than the original estimates. Since there is no clear benefit from adding this extra parameter and the theoretical justification as a model for nonlinear effects is weak, its inclusion in our main-stream analysis does not seem to be justified. Still, in our total error-balance, we consider a systematic error of $15 \%$ due to non-linear effects. \subsection{Comparison to the PS from Mark III} \label{subsec:mark3} A similar likelihood analysis (though with errors fixed a priori) has been recently applied by Zaroubi {\it et al.\ } (1997) to the Mark~III catalog of peculiar velocities. (Willick {\it et al.\ } 1995; 1996; 1997a). It is interesting to investigate whether the recovered power spectra are consistent with each other, given the respective uncertainties. This is intriguing because there are certain differences in the velocity fields as reconstructed from the two samples, especially in the bulk flows both in the very local neighborhood and of outer shells ({\it e.g.}, da Costa {\it et al.\ } 1996; 1998; Dekel 1998; Dekel {\it et al.\ } 1998; Giovanelli {\it et al.\ } 1998a, 1998b). \begin{figure}[tbp] {\special{psfile="f11.ps" angle= 0 hscale=90 vscale=90 hoffset=-35 voffset=-554}} \vspace{14.5 cm} \caption{\capt Comparison of SFI PS results to Mark III for the tilted \lcdm test-case. Top left panel shows ${\rm ln}\L$ contours for the SFI data, top right panel shows the contours for the Mark III data. The best-fit parameters for SFI and Mark III are marked, on both, by `S' and `M' respectively. The lower panel shows the maximum likelihood PS corresponding to SFI (solid) and Mark III (dashed). The 3 solid dots mark the PS calculated from Mark III by Kolatt and Dekel (1997), together with their estimated $1\sigma$ error-bar. The shaded region is the SFI likelihood $90\%$ confidence region. } \label{fig:mark3} \end{figure} Figure~\ref{fig:mark3} presents a comparison of the likelihood analysis for the two samples using our representative tilted \lcdm family of models. Shown once again are the likelihood contours for the SFI data, together with the corresponding plot for the Mark III data. The high-likelihood ridge is similar for both samples. While the Mark~III result slightly favors higher values of $\Omega$ and lower values of $n$, the differences are along the ridge of maximum likelihood and are therefore hardly significant. The contours are slightly more concentrated for the Mark III catalog because it consists of more galaxies. The best-fit PS for the two catalogs are shown in the bottom panel of Figure~\ref{fig:mark3}, on top of the shaded area which marks the $90\%$ confidence region for SFI. The resultant power spectra are consistent within the errors, and they agree particularly well on intermediate scales, where the data provides the most meaningful constraints. The corresponding best values for $\sigma_8 \Omega^{0.6}$ are 0.81 and 0.85 for SFI and Mark~III respectively. Similar results are obtained when comparing likelihood analysis of the two catalogs using the other PS models. It is worth noting here that the systematics discussed in the previous sections, with regard to the SFI analysis, are found to affect the Mark~III likelihood analysis in a similar way. The PS computed by Kolatt \& Dekel (1997) from the Mark~III smoothed density field recovered by POTENT is also displayed on the figure (as three symbols with error bars). The agreement of the SFI result with this independent calculation of the Mark~III PS is good. A recent comparison of Mark III with IRAS 1.2 Jy (Willick \& Strauss 1998) suggests an alternative zero-point calibration for one of the Mark III datasets. We have applied our likelihood analysis to the Mark III data revised accordingly, and found negligible changes in the resulting power spectrum and cosmological parameters, smaller than the uncertainties due to global zero-point discussed in \S~\ref{subsec:0p}. The close agreement between the mass power spectra derived from the two datasets indicates that the results presented here are quite robust and are unlikely to arise form specific peculiarities of either of the two samples. This does not preclude possible differences that are not picked up by the specific statistic used -- in our case, the mass PS. In particular, the difference in the two bulk flows, which is known to exist, is not reflected in the power spectra. This is because the wavenumbers corresponding to the bulk velocity are smaller than the $k$ range that dominates the fit in our current analysis. \section{CONCLUSION} \label{sec:concl} We used a linear maximum-likelihood method to measure the mass-density power spectrum from the SFI catalog of peculiar velocities, and to determine the cosmological parameters for families of physical CDM models with or without COBE normalization. We have corrected for biases introduced by the non-trivial selection procedure of the SFI catalog using a new semi-analytic procedure. We have verified that the results are quite insensitive to the detailed way by which we implement this bias correction. This approach allows also to refine the distance errors estimates. Our new version of likelihood analysis enabled us to independently verify the error estimates of the SFI catalog to within an uncertainty as small as $5\%$ of the error, which we regard as very encouraging. Since the errors affect the PS in a systematic way, this independent confirmation adds significantly to our confidence in the results. The general result for all the models examined here is that the power spectrum at $k= 0.1 \, h\, {\rm Mpc^{-1}}$ is $P(k) \Omega^{1.2} = (4.4\pm1.7)\times 10^3 \, ( h^{-1} {\rm Mpc})^3$, and that $\sigma_8 \Omega^{0.6} = 0.82 \pm 0.12$. These results are obtained by the peculiar-velocity data independent of the specific shape assumed for the PS, and are consistent with the result of the $\Gamma$ model independent of the COBE normalization. The random errors quoted are $90\%$ confidence level and they include small variations due to the choice of model for the PS within the families of models tried here. For the general family of COBE-normalized CDM models, we find a high-likelihood ridge in the $\Omega-n-h$ parameter space, which can be crudely approximated by $\Omega\, n^{\nu}\, {h_{60}}^{\mu} = 0.62 \pm 0.15$, where for \lcdm $\mu=1.3$ and $\nu=2.0,3.9$, without and with tensor fluctuation respectively. For \ocdm, without tensor fluctuations, the powers are $\mu=0.9$ and $\nu=1.4$. Again, the error quoted is the formal $90\%$ uncertainty including the model variations. Thus, for $h=0.6$, the maximum-likelihood value of $\Omega$ ranges between 0.6 and unity while $n$ varies between 1 and 0.8 respectively. Without a tilt, values of $\Omega$ as low as 0.5 are allowed within the $90\%$ confidence limit. Our tests using mock catalogs based on an $N$-body simulation that mimics our cosmological neighborhood indicate that the systematic errors in our results are relatively small. In particular, the nonlinear effects in the mock catalogs are found to be negligible. This is indeed expected because the quantity we actually measure is the velocity power spectrum in the mildly-nonlinear regime, which we have demonstrated to be reasonably approximated by linear theory. An ad-hoc test for nonlinear (multi-streaming) effects in the data themselves indicated that they may work to reduce the values of the cosmological parameters given above, but that this effect is not larger than $\sim 15\%$. In order to refine our estimates of the systematic effects even further, we intend to repeat the current analysis using a proper nonlinear scheme, and to repeat the tests of the method using simulations of higher resolution which are in preparation. We thus estimate the total systematic uncertainty to be of order $\sim 15\%$, namely comparable in size to the random errors. Therefore, to be on the safe side when comparing our results to other results, we recommend as a rule of thumb multiplying the quoted errors by a factor of $\sim 2$. As yet another word of caution, it is worth recalling that our analysis is heavily weighted by the galaxies at relatively small distances, because the data is weighted by the inverse squared of the distance errors. This means that the result is sensitive to the data and error estimate of the inner galaxies. It is possible in principle that a source of distance error which operates preferentially at small distances has somehow escaped our attention and is not properly modeled by our error model. To test the effect of such a possibility, we have repeated the analysis after pruning all galaxies with distances smaller than a given distance. When pruning inside $15\,{\rm h^{-1}Mpc}$ ($3\%$ of the data) we obtain for the most likely value $\sigma_8\Omega^{0.6} =0.85$ instead of the original result of $\sigma_8\Omega^{0.6} =0.81$ when using all the data (still with our standard pruning based on linewidth). When pruning inside $25\,{\rm h^{-1}Mpc}$ ($17\%$ of the galaxies), we obtain instead $\sigma_8\Omega^{0.6} =0.71$. It is encouraging to find that these variations are within the $90\%$ likelihood contours of the different cases in the $\Omega-n$ plane, but this is yet another potential source of uncertainty to bear in mind. A systematic trend does seem to show up when we eliminate as much as the whole inner half of the data (inside a distance of $46\,{\rm h^{-1}Mpc}$, or with linewidth smaller than $2.48$); the outer data, when analyzed by themselves, indicate a significantly lower PS then the inner data. This effect is not reproduced in the mock catalogs and is therefore not likely to represent a general fault in the method. Possible explanations for this effect are larger uncertainty in our estimate of random and systematic errors at large distances, differences between the assumed and the true TF relation, or a true PS with a different shape than our models. It may also be due to a real difference between the density fields in the two halves (that is somehow not properly reproduced in the simulation) or to a systematic dependence of velocity bias on galaxy properties. We carried out a number of tests in which we added to the likelihood analysis ad-hoc free parameters which allow more flexibility in the distance dependence. These include variations in the TF parameters and in the errors as a function of linewidth. Our tests indeed led to some improvement in the agreement between the two subsamples, but, being only preliminary, they did not yield so far a firm conclusion as for the dominant source of the effect and the optimal way to deal with it. Since the variations introduced preferentially affect the peculiar velocities of galaxies at large distances, which typically have large errors and therefore contribute only little to the likelihood procedure, they do not affect significantly the resultant power spectrum from the full sample. We therefore conclude that our current results are robust, and defer a more thorough investigation of this trend to a future analysis. The recovered mass power spectrum, and the constraints on the cosmological parameters obtained here, are consistent with the results of a similar analysis applied to the Mark~III catalog of peculiar velocity. This is despite the fact that these two catalogs seem to differ in some of their other properties, such as the large-scale bulk velocity. Indeed, the bulk velocity is not expected to contribute to the density on smaller scales. There is also an apparent disagreement between the results obtained from peculiar velocities of clusters (Borgani {\it et al.\ } 1997) and our result for the SFI field galaxies. As mentioned in the Introduction, our dynamical result of $\tilde\sigma_8 \equiv \sigma_8 \Omega^{0.6} \simeq 0.8\pm 0.2$ may be crudely compared to estimates of the $\beta$ parameter obtained when comparing the same SFI data to a redshift survey of galaxies. da Costa {\it et al.\ } (1998) find $\beta=0.6\pm 0.1$ when comparing the SFI peculiar velocities to the velocities predicted by the IRAS 1.2 Jy redshift survey, assuming linear biasing. A similar value was obtained from Mark III when the comparison was done via velocities (Davis {\it et al.\ } 1996). With $\sigma_{8\rm g} \simeq 0.7$ for IRAS galaxies, the predicted $\beta$ from our current constraint on $\tilde\sigma_8$, via $\beta=\tilde\sigma_8 / \sigma_{8\rm g}$, is significantly closer to unity than to 0.6 (compare also to Kolatt \& Dekel 1997, Fig. 6). The residuals between the measured peculiar velocities and the IRAS predictions, for the best-fit $\beta$ value, were found in this comparison (da Costa {\it et al.\ } 1998) to be significantly higher than the errors as originally estimated for the SFI data (based on the scatter observed in the SCI cluster sample) combined with the errors estimated for the IRAS data. One possibility is that the the IRAS model fails to predict some of the peculiar velocities that exist in the SFI data, e.g., because the distribution of galaxies is not properly approximated by a simple, linear, scale-independent and deterministic biasing relation ({\it e.g.}, Dekel \& Lahav 1998). In that case, the interpretation of the value of $\beta$ determined from fitting the IRAS predictions to the velocity data is not clear. On the other hand, it is also possible that the errors in SFI are indeed larger than originally estimated. Such larger errors would accordingly reduce the PS amplitude, in particular the value of $\tilde\sigma_8$, as estimated in the current paper. However, such an effect should have been detected by investigating the global biasing properties of the sample. It would also be hard to understand why our likelihood analysis does prefer errors very similar to the original estimates. Although we are fairly convinced that nonlinear effects in the current analysis are confined to the level of $\leq 15\%$, it will be worth making an extra effort to improve the accuracy in a future paper. \acknowledgments{We thank George Blumenthal, Stefano Borgani, Gerard Lemson, Adi Nusser, and Michael Strauss for stimulating discussions. We are grateful to the ESO Visitors fund for supporting visits to Garching by AD, GW, IZ, MPH, and RG. This research was supported in part by an ESO DGDF grant to WF, the US-Israel Binational Science Foundation grant 95-00330 and the Israel Science Foundation grant 950/95 to AD, IZ and AE, the DOE and the NASA grant NAG 5-7092 at Fermilab to IZ, NSF grants AST94-20505 and AST96-17069 to RG, AST95-28860 to MPH, and AST93-47714 to GW.}
\section{Introduction} The landscape paradigm \cite{landscape} is a very useful point of view for the study of glassy systems. The detailed analysis of the free energy or potential energy surfaces allows us to get insight into the rich phenomenology exhibited by glass-forming liquids in the supercooled phase, around the glass transition region and in the low temperature glassy regime. While the investigation of the free energy surface is a very hard task starting from the microscopic description of the system, since the landscape details are very strongly temperature dependent, the description of the potential energy landscape is a much more tractable problem, and is a good starting point to investigate properties at not-too-high temperatures. The trajectory of the representative point of the system in the configurational phase space can be viewed as a path in the multidimensional potential energy surface. The dynamics is strongly influenced by the topography of the landscape: local minima, barriers height, attraction basins and further topological features. In recent years it has been show that the details of the potential energy surface are of great importance in determining the properties of many systems exhibiting glassy behavior, like glass-forming liquids, protein folding, atomic cluster or evolutionary biological models. In this paper we numerically investigate the low temperature dynamical properties of a simple system, i.e. a monoatomic Lennard-Jones system, through an analysis of its multidimensional potential energy surface and a simple model for the loe temperature dynamics. In the first part (Section II), by studying small sample size, we give a comprehensive description of the potential energy landscape of the systems: minima, barriers, reaction paths, saddle points, and determine statistical distributions and cross-correlations among the analyzed quantities. In the second part (Section III) we use this information to set up a simple model for the study of the long-time relaxation dynamics of the system; the model consists of a connected network of potential energy minima with a jump dynamics described by an appropriate master equation. This allows us to get information on the behavior of the system for times long enough that a direct Molecular Dynamics (MD) simulation is not feasible \cite{lettera}. After a static (thermodynamical) test of the model, we determine the dynamical equilibrium properties and also the off-equilibrium ones, and discuss the results. In the last Section we report the conclusions. \section{Potential energy landscape} We numerically investigate the topology of the potential energy hypersurface of a Lennard-Jones $6-12$ system of $N$ interacting particles in a cubic box with periodic boundary conditions. The pair potential is: \begin{equation} V_{LJ}(r)=4\epsilon \ \left[ \left( \frac \sigma r\right) ^{12}-\left( \frac \sigma r\right) ^6\right] \ , \end{equation} with $r$ the Euclidean distance between two particles. The physical parameters $\sigma $ and $\epsilon $ are choose to describe an Argon system: $\sigma =0.3405$ nm and $\epsilon /K_B=125.2$ K ($K_B$ is the Boltzmann constant). In order to have an exhaustive description of the energy landscape, we investigate small size systems, with a number of particles $N<30$. As we shall see later, such small systems can nevertheless exhibit complex enough behavior. Due to the small size of the system, the range of interactions among the particles is of the same order as the box length. It is then much more appropriate to use a multi-image method instead of the usual minimum image method \cite{ALLEN}, in which each particle interacts only with the nearest image (generated by the periodic boundary conditions) of all other particles. In our case many images contribute, and it is necessary to consider them in the calculation of the interactions. The choice we made introduces an angular dependence in the pair potential which, however, is negligible with respect to the radial one. The simulated density is $\rho =4.2\cdot 10^{-2}$ mol/cm$^3$, which was obtained by Demichelis et al. \cite{DEMICHE} as the smallest value at which a Lennard-Jones system of $N=864$ Argon particles is stable in the glassy state after a rapid quench from high temperature. \subsection{Minima} As a first step we search for the potential energy minima. We used a modified conjugate gradient method starting from high temperature configurations obtained during a MD simulation. In this way we find the so called ``inherent configurations'', corresponding to local minima. New minima are identified by their potential energy values. We analyzed systems with $N=11\div 29$ and, for each $N$, we stopped the search when the rate at which new low energy minima are found (for example in lowest first third of the full energy range) is smaller than a given number (about $10^{-4}$). In this way we were able to obtain a good classification of the low energy minima. The number of found minima, ${\cal N}$, does not show a clear and well defined dependence on $N$, contrary to the case of small clusters \cite{DALDOSS}, so that it is not possible to give an estimate of the coefficient $\alpha $ in the expected exponential growth: ${\cal N}\propto \exp (\alpha N)$ (for clusters $\alpha \sim 1$). We observe indeed for some values of $N$ a strong tendency of the system to fall always in the same minima. In this case an exhaustive research of the inherent structures is a very hard numerical task. this explains the not clear dependence found. Once most minima are classified, we analyze their features: energy, static structure factor, curvature and distances. \subsubsection{Energy} The first important information about an inherent configuration is its potential energy value. From the energy distribution it is sometimes possible to recognize the crystalline-like configurations. Usually one finds an evident gap between the lowest energy minima, the crystalline-like ones, and the other ones with higher energies, as can be seen in Fig. $1$ where we report the distribution of minima energies per particle for $N=29$. In other cases the situation is not so clear and it is useful to use a different method to characterize the nature of an inherent configuration; the most useful quantity in determining the spatial order structure of a minimum is its static structure factor, which allows (imperfect) crystal-like configuration also of high energy to be identified. \subsubsection{Static structure factor} In order to classify the spatial distribution of particle in a given minimum of the potential energy we use the static structure factor: \begin{equation} {\hat S}({\vec q})=\frac 1N|\sum_j\exp (i{\vec q}\cdot {\vec r}_j)|^2\ . \end{equation} Due to the finite size of the system the allowed ${\vec q}$ vectors are of the form ${\vec q}=(2\pi /L)\ {\vec n}$, with ${\vec n}=(n_x,n_y,n_z)$ an integer vector. We define the quantity \begin{equation} S(q)=\frac 1{n_q}\sum_{\vec q \in \{q,\Delta q\}}{\hat S}({\vec q})\ , \end{equation} where $\sum_{\vec q \in \{q,\Delta q\}}$ is a sum over the $n_q$ vectors with modulus within $q \pm \Delta q$. For a pure crystalline configuration of $N$ particles $S(q)$ consists of Bragg peaks, and its value at the peaks is $S_{max}=N$. For amorphous configurations $S(q)$ does not present a well defined peak structure and the highest value is $S_{max}\sim 2\div 3$, obtained for a $q$ value of the order of the inverse mean distance of two near particles. For small sized systems, like those analyzed here, there are intermediate situations and we use the criterion $S_{max}<N/2$ to determine the amorphous nature of an inherent structure. \subsubsection{Curvature} Another important property of a minimum is its overall curvature $c$, defined as the determinant of the Hessian $\Phi^{^{\prime \prime }}$ of the potential energy function $\Phi$: \begin{equation} c=\det (\Phi^{^{\prime \prime }})\ , \end{equation} The eigenvalues of the Hessian matrix are proportional to the squared vibrational eigenfrequencies. In the inset of Fig. 1 we show the distribution of \begin{equation} \gamma=\frac 1{3N-3}\ {\rm Log}_{10}\ (c/m)^{1/2}\ , \end{equation} where $m$ is the mass of the particles (we use the Argon's value $m=40$ amu). The quantity $\gamma$ is thus proportional to the sum of the logarithms of the frequencies of normal vibrational modes \begin{equation} \gamma=\frac 1{3N-3}\ \sum_\alpha \ \ln \ \omega _\alpha\ , \end{equation} with $\alpha =1,...,N-3$ (the three zero frequencies corresponding to rigid translations have been eliminated from the sum). As can be seen in Fig. 1, the highest $\gamma$ values correspond to the minima with lowest energy, i.e. the crystalline-like minima which are narrower and deeper than the other packing structures (see also Sec. II-C). \subsubsection{Stress tensor} For each minimum we determined the off-diagonal part of the microscopic stress tensor: \begin{equation} \left\{ \begin{array}{rcl} \displaystyle \sigma ^{zx} & = & \displaystyle -\sum_{i<j}V_{LJ}^{^{\prime }}(r_{ij})\frac{z_{ij}x_{ij}}{r_{ij}}\ , \\ & & \\ \displaystyle \sigma ^{xy} & = & \displaystyle -\sum_{i<j}V_{LJ}^{^{\prime }}(r_{ij})\frac{x_{ij}y_{ij}}{r_{ij}}\ , \\ & & \\ \displaystyle \sigma ^{yz} & = & \displaystyle -\sum_{i<j}V_{LJ}^{^{\prime }}(r_{ij})\frac{y_{ij}z_{ij}}{r_{ij}}\ , \end{array} \right. \label{tsfo1} \end{equation} where $x_{ij},y_{ij},z_{ij}$ are the components of ${\vec r}_{ij}= {\vec r}_i-{\vec r}_j$; these quantities will be useful in determining the shear viscosity. The form of the stress tensor we use is the $q=0$ extrapolation of the $q$-dependent expression \cite{BALUCANI} \begin{eqnarray} \sigma _{\alpha ,\beta }({\vec q}) &=&\sum_i\left[ mv_{i,\alpha }v_{i,\beta }-\sum_{i<j}\frac{r_{ij,\alpha }r_{ij,\beta }}{r_{ij}^2}P_q(r_{ij})\right] \cdot \nonumber \label{tsfo2} \\ &&\cdot \exp (i{\vec q}\cdot {\vec r}_i)\ , \end{eqnarray} with \begin{equation} P_q(r)=rV_{LJ}^{^{\prime }}(r)\frac{1-\exp (i{\vec q}\cdot {\vec r})}{i{\vec q}\cdot {\vec r}}\ . \end{equation} The index $i$ and $j$ refer to particles, while $\alpha $ and $\beta $ label the spatial axes $x,y,z$. In (\ref{tsfo1}) we have omitted the kinetic term, not well defined in an inherent configuration; our hypothesis is that also this truncated form is able to well describe the relaxation processes. \subsubsection{Distances} We now turn to the relationships among the minima; in particular we determined the mutual distances in the $3N$-dimensional configuration space; in this regard it is important to take into account all the symmetry operations of the problem. Indicating with ${\underline{r}}_a=({\vec r}_1^a,..., {\vec r}_N^a)$ the $3N$ coordinates of the particles in the minimum $a$, we define the distances $d_{ab}$ between minima $a$ e $b$: \begin{equation} d_{ab}=\min_{{\small {T,R,\pi }}}(|{\underline{r}}_a-{\underline{r}}_b|)\ , \label{dista} \end{equation} where the minimization is made with respect to the continuous translations ($T$), discrete rotations and reflections ($R$), and permutations ($\pi $) of the particles. The minimization over the continuous translations ($T$) is done by putting, sequentially, each particle of $a$ in the same place as each one of $b$. The minimization over the rotations and reflections ($R$) is carried out by considering the $48$ symmetry operations of the cubic group; the minimization over particle permutation ($\pi $) is apparently an hard task to solve, as one should consider all the $N!$ possible configurations in a direct calculation, but this is not the case actually. The problem is of polynomial type, i.e. the time needed to find the optimum solution grows as a polynomial function of the size, $N$, of the system (on the other hand non-polynomial problems require an exponential computational time, i.e. the travel salesman problem). The optimization problem we need to solve is a bipartite matching problem, which can be done in very short computational time by using an appropriate algorithm. \subsection{Barriers} A very important topological quantity in determining the dynamical behavior of the system is the energy barrier experienced by the system in traveling from one minimum to another. At first sight it might seem that it is accurate enough to evaluate the barrier along the straight path joining two minima in the $3N$-dimensional space. However, as evidenced by Demichelis et al. \cite{DEMICHE}, this produces in most cases much higher barriers than an evaluation along the least action path, indicating that the straight path approximation is often not good. We have then determined the least action path for each pair of minima $a$ and $b$, which is defined as the path that minimizes the action functional \begin{equation} S[\ell ]=\int_\ell ds\sqrt{2m[\Phi ({\underline{r}}(s))-\Phi _0]}\ , \end{equation} where $\ell $ is a generic path between the minima, $s$ is the curvilinear coordinate and $\Phi _0=\min \{\Phi _a,\Phi _b\}$. This functional problem is simplified by dividing the path into a finite number of intervals (typically we use $n=16$ intervals) and by minimizing the action function with respect to the extreme of the $n$ segments constrained to move in hyperplanes perpendicular to the straight path. The highest potential-energy value along the least action path determines the barrier height and identifies the saddle point. We have only analyzed the system with $N\leq 17$, due to the very long computational times needed in the cases with $N>17$ where there are too many pairs of minima to take into account. We report the results for the largest system analyzed, $N=17$ with ${\cal N}=38$ minima. In Fig. $2$ we show an example of the potential energy profile along the straight path (dashed line) and along the least action path (full line) between two minima. It is evident that the energy barrier is significantly lower in the latter case, although the two paths are not very distant in configuration space. Similar results are obtained for the other analyzed paths. Sometimes it happens that two minima are not directly connected, in the sense that the least action path joining them crosses a third minimum, and a non trivial connectivity among the minima emerges (in the analyzed system we find that each minimum on an average is directly connected to $20$ other minima). For each saddle point along the least action path we determine the main properties: energy, curvature, down ''frequency''. In Fig. $3$ is reported the energy distribution of the barriers $\Delta\Phi_{bar.}$. The curvature is defined as the absolute value of the determinant of the Hessian of the potential energy evaluated at the saddle: \begin{equation} c_{sad.}=|\det \Phi _{sad.}^{^{\prime \prime }}|\ , \end{equation} In the inset of Fig. $3$ we show the distribution of the quantity \begin{equation} \gamma _{sad.}=\frac 1{3N-3}\ {\rm Log}_{10}\ (c_{sad.}/m)^{1/2}\ . \end{equation} The down ``frequency'' ${\tilde \omega }_{sad.}$ is defined as the square root of the absolute value of the down curvature along the least action path: \begin{equation} {\tilde \omega }_{sad.}^2=-\frac{{\underline{v}}\ \Phi _{sad.}^{^{\prime \prime }}\ {\underline{v}}}{|{\underline{v}}|^2}\ , \end{equation} where ${\underline{v}}$ is the tangent vector to the least action path at the saddle point. \subsection{Correlations} In order to have a full statistical description of the potential energy landscape, it is useful to investigate, besides the distributions of the different quantities, also their cross-correlations. We have then determined the linear correlation coefficient \begin{equation} r(x,y)=\frac{\sum_i(x_i-{\overline{x}})(y_i-{\overline{y}})}{[\sum_i(x_i- {\overline{x}})^2\ \sum_j(y_j-{\overline{y}})^2]^{1/2}}\ \end{equation} for all the measured quantities $x$ and $y$. In Table I are reported the values obtained, together with the $\log -\log $ correlations, in order to evidence possible power laws. In the first column we report the number of particles of the system analyzed (only for $N=17$ we have determined the least action paths and all the related quantities). It emerges that, substantially, energy difference and distance among the minima are not correlated, indicating that the topological structure of the inherent configurations is not energy-correlated. A weak correlation is observed between energy and curvature at stationary points (minima and saddle points). In Fig. $4$ we show the cross-correlation between the energies of the minima and their curvatures. An interesting correlation is observed between barrier energies $\Delta\Phi_{bar.}$ and distances among minima $d_{ab}$ (Fig. $5$), with a nearly linear correlation in double $\log $ scale (line in the figure). We conclude the analysis of the energy landscape by determining the entropy ratio $R$, defined as the ratio between the curvature of saddle points and that of the related minima \begin{equation} R=\frac{|\det \Phi _{sad.}^{^{\prime \prime }}|}{\det \Phi _{min.}^{^{\prime \prime }}}\ . \end{equation} This quantity gives a quantitative measure of the ability of the system in finding the right path to reach another minimum. If $R\sim 1$ there are no entropic hindrances, while if $R\gg 1$ these effects become relevant, as the narrowness (higher value of the curvature) at the saddle makes the least action path toward that specific minimum unfavorable with respect to other escape route. For all the minimum-saddle-minimum triplets we have evaluated $R$; the majority of the values is in the range $10^{-2}\div 10$, in qualitative agreement with the results found in Lennard-Jones clusters \cite{DALDOSS}. \section{Model for the dynamics} The investigation of the properties of glass-former liquids at the level of the energy landscape, allows us to introduce some approximations in the dynamics of the system. We define a simplified model which is able to capture the long time behavior of the system, and which consists of a connected network of potential energy minima with a jump dynamics among them described by an appropriate master equation \cite{lettera}. The basic idea is quite simple. A glass is represented by a configurational point confined in a very small region of the accessible phase space and in the zero temperature limit (neglecting quantum effects) all the atoms are frozen in well defined positions, corresponding to some mechanically metastable state. When the temperature is raised jumps among different mechanically stable positions become possible. At finite and not-too-high temperature we assume the dynamically relevant processes are the following: a short time dynamics dominates by small vibrations around stable positions (this dynamics can be described within the harmonic approximation by diagonalizing the dynamical matrix), and a long time dynamics consisting of collective (involving many atoms) jumps among different stable positions. The main hypothesis we make is that there is a substantially clear separation of time scales between the two dynamical processes. This characterization of the dynamics is a good approximation at not-too-high temperature. By increasing the temperature, anharmonic effects become relevant to the vibration around the local minimum and, moreover, a clear time scale separation between fast vibrational and slow jumps dynamics is no longer possible. In a recent work \cite{SCHRO} the validity of this hypothesis has been verified in a Lennard-Jones binary mixture, giving a very good agreement with a direct MD investigation. To sum up, our model, which is expected to capture the physics of the system at low temperature, is based on two main hypotheses: \begin{enumerate} \item clear cut difference between vibrational dynamics at short time and dynamics of collective jumps at long time; \item description of the long time dynamics through a master equation, with the transition rates that depend on the topological properties of the potential surface. \end{enumerate} The main advantages of the introduced model with respect to the usual MD computations are: \begin{itemize} \item we can avoid in a simple way the crystallization process, that always takes place in one component LJ systems, as we do not consider the crystalline minima in setting up the network. \item we can study in a direct way the low temperature properties, where usually the very long relaxation times require very long computational time. In MD the computational times are proportional to the physical times, while in the model introduced here the computational times are those needed to find the eigenvalues and eigenvectors of the transition matrix, independent of temperature; \item it is possible to evidence the relationships between the energy landscape and the behavior of the system. \end{itemize} To be more specific, the model is a connected network of potential energy minima and the master equation governing the jumps dynamics is: \begin{equation} \frac{dp_a}{dt}(t;b,t_0)=\sum_c\ W_{ac}\ p_c(t;b,t_0)\ , \label{me} \end{equation} where $p_a(t;b,t_0)$ is the probability that the system is at minimum $a$ at time $t$, if it was at minimum $b$ at time $t_0$. The off-diagonal elements of the matrix $W$ are the transition rates. The diagonal elements are fixed by the condition \begin{equation} \sum_aW_{ac}=0\ . \end{equation} In order to obtain an asymptotic behavior that reproduces the right Boltzmann weight the occupation probability must satisfy: \begin{equation} \lim_{t\rightarrow \infty }p_a(t;b,t_0)=p_a^0\equiv \frac 1{{\cal Z}}(\det \Phi _a^{^{\prime \prime }})^{-1/2}\ \exp (-\beta \Phi _a)\ , \end{equation} (${\cal Z}$ is such that $\sum_ap_a^0=1$ and the pre-exponential factor follows from the harmonic vibration in each minimum), and the transition matrix $W$ must satisfy the detailed balance relation: \begin{equation} W_{ab}\ p_b^0=W_{ba}\ p_a^0\ . \end{equation} The solution of the master equation is easily expressed in terms of eigenvalues $\lambda _n$ and eigenvectors $\alpha _a^{(n)}$ ($n=1,...,M$, with $M$ matrix dimension) of $W$: \begin{equation} p_a(t;b,t_0)=(p_b^0)^{-1}\ \sum_n\ \alpha _a^{(n)}\ \alpha _b^{(n)}\ \exp [\lambda _n(t-t_0)]\ . \label{me2} \end{equation} In the numerical calculus it is more convenient to express the solutions in terms$_{}$ of the eigenvectors of a new symmetric matrix $w_{ab}=W_{ab}\ (p_b^0/p_a^0)^{1/2}$ (whose eigenvalues coincide with those of $W$): \begin{equation} p_a(t;b,t_0)=(p_a^0/p_b^0)^{1/2}\ \sum_n\ e_a^{(n)}\ e_b^{(n)}\ \exp [\lambda _n(t-t_0)]\ . \label{me3} \end{equation} The model is well defined once we give an appropriate form to the transition matrix $W$. In order to determine the transition rates let us analyze the problem of escape from a metastable state; a useful point of view for systems with many degree of freedom is the description in term of a few relevant coordinates. This reduction is possible whenever there are few reaction coordinates with characteristic evolutionary times longer than those of the other degrees of freedom, which act as effective terms on the relevant coordinates, i.e. like noise and viscous terms. We suppose this is the case of our system whenever the temperature is not too high (the analysis of reaction paths made by Demichelis et al. \cite{DEMICHE} supports this hypothesis). Handling the problem as a Markovian-Brownian $d$-dimensional motion in the overdamped friction regime we obtain the form \cite{RISKEN} \begin{equation} W_{ab}=\frac{\tilde \omega _{sad.}^2}\mu \left[ \frac{\det \Phi _b^{^{\prime \prime }}}{|\det \Phi _{sad.}^{^{\prime \prime }}|}\right] ^{1/2}\ \exp \left[ -\frac{\Phi _{sad.}-\Phi _b}{K_BT}\right] \ , \label{wab} \end{equation} where $\tilde \omega _{sad.}$ is the down ``frequency'' at the saddle point and $\mu $ is a friction constant that determines the time scale (its value is fixed by a comparison with MD in the allowed temperature region). All the characteristics of the model (properties of the connected network and parameters in the transition rates) are inferred from the computed properties of the potential energy landscape. We use the values of the $N=29 $ system to determine local minima properties (energy, curvature, stress tensor) and those of $N=17$ system to determine connectivity properties (energy and curvature of saddle points, distances and connectivity among the minima). The values are extracted from the distribution found in the previous section in the following way: \begin{enumerate} \item we extract $M$ energy values of the minima from the distribution of $N=29$ system (we exclude the crystalline-like configurations); \item we assign to each minimum a value of curvature $c_a=\det \Phi _a^{^{\prime \prime }}$ extracted from a bivariate distribution, thanks to the cross-correlation between energy and curvature; a stress tensor value is also extracted for each minimum; \item for each minimum we randomly (in the analysis of the energy landscape we have found no correlation between energies and distances among minima) extract $20$ minima connected to it, as obtained on an average for the system $N=17$; \item we define a connection matrix $\kappa _{ab}$, containing the minimum steps, i.e. the number of minima crossed, necessary to go from $a$ to $b$; the distance matrix $d_{ab}$ is $\kappa _{ab}$ times the value extracted from the distribution of the distances among connected minima for $N=17$; \item for each pair of directly connected minima we determine the energy barriers $\Delta\Phi_{bar.}$ from the value of the distance $d_{ab}$: $\Delta\Phi_{bar.}=A\ d^\alpha $ ($A\simeq 10^5$ and $\alpha \simeq 3.7$, as determined for $N=17$ system, Fig. 5); \item we assign a curvature value $c_{sad.}=|\det \Phi _{sad.}^{^{\prime \prime }}|$ and a down ``frequency'' $\tilde \omega _{sad.}$ to each saddle point, from bivariate distribution. \end{enumerate} We obtain in this way a set of parameters that describe the model. In order to have a good statistic description we considered different extractions of the parameters and the measured quantities were obtained by averaging over the extractions. \subsection{Test} Before studying the dynamical properties of the model, we concentrate on the static behavior obtained as asymptotic solution of the master equation. In this static regime we can determine, in a statistical mechanical approach, the configurational partition function \begin{equation} {\cal Z}(\beta )=\int d^{3N}r\ \exp [-\beta \Phi ({\vec r}_1,...,{\vec r}_N)] \ , \end{equation} By using the approximation based on the hypothesis of short time local harmonic vibrations around a minimum, and long time collective jumps among different minima, we obtain \begin{equation} {\cal Z}(\beta )\sim \sum_a{\cal Z}_a^{(harm.)}(\beta )\ \exp [-\beta \Phi _a]\ , \end{equation} where $a$ labels the minima and ${\cal Z}_a^{(harm.)}$ is the contribution of harmonic vibrations around minimum $a$. This form of the configurational partition function emerges in the model as the exact infinite-time limit. The harmonic term is easy to calculate, being a $3N$-dimensional Gaussian integral: \begin{eqnarray} {\cal Z}_a^{(harm.)}(\beta ) &=&\int d{\underline{r}}\exp \left[- \frac \beta 2\ {\underline{r}}\ \Phi _a^{^{\prime \prime }}\ {\underline{r}}\right] \nonumber \\ \ &=&(2\pi)^{3N/2}\ \beta ^{-3N/2}\ (\det \Phi _a^{^{\prime \prime }})^{-1/2}\ , \end{eqnarray} where ${\underline{r}}=(r_1,...,r_{3N})$, and ${\underline{r}}\ \Phi _a^{^{\prime \prime }}\ {\underline{r}}=\sum_{l,m}r_l\ (\Phi _a^{^{\prime \prime }})_{lm}\ r_m$. We then obtain the approximated partition function as: \begin{equation} {\cal Z}(\beta )\sim c\ \beta ^{-3N/2}\ \sum_a\ (\det \Phi _a^{^{\prime \prime }})^{-1/2}\ \exp (-\beta \Phi _a)\ , \label{zapp} \end{equation} from which the thermodynamical quantities can be derived, for example for the energy: $E(\beta )=- \partial _\beta \ $ln\ ${\cal Z}$. To check the reliability of the model we compare the quantities calculated from (\ref{zapp}) with those obtained through MD computation. In Fig. $6$ we show the potential energy as obtained from the model (lines) and from MD (circles). The MD data are obtained in the following way. Starting from high temperature we quench rapidly the system to low temperature, entering in a glassy state; we then increase the temperature up to liquid phase (open circles). The system is subsequently slowly cooled, entering in the supercooled regime ($100-70$ K) and obtaining at the end the crystal through a first order transition (full circles). The lines represent the energies determined from the model by taking into account all the minima (dotted line) and only the glassy ones (full line). A good quantitative agreement is obtained between MD and the model as far as the temperature is lower than about $150$ K, a temperature in the liquid phase well above the melting point ($T_m\sim 80$ K). This result supports the correctness of the approximation of local-vibration / collective-jumps in the description of a glass-former at not-too-high temperature. This static test is a good starting point to extend the analysis to the dynamical regime. \hfill \subsection{Equilibrium properties} We now determine the dynamical equilibrium properties of the model. We denote with ${\cal O}({\underline{r}}(t),{\underline{r}}(0))$ a generic observable which depends on collective coordinates ${\underline{r}}$ at time $t$ and at initial time $t=0$. We define the statistical average value of ${\cal O}$ in the model as \begin{equation} <{\cal O}(t)>=\sum_bp_b^0\ \sum_a{\cal O}_{ab}\ p_a(t;b,0)\ , \label{oss1} \end{equation} where ${\cal O}_{ab}$ is the value of ${\cal O}$ evaluated at the minimum configurations $a$ and $b$: ${\cal O}_{ab}={\cal O}({\underline{r}}_a, {\underline{r}}_b)$. In terms of the eigenvalues and eigenvectors of the transition matrix $W$ we have: \begin{equation} <{\cal O}(t)>=\sum_n\exp (\lambda _nt)\ \sum_{a,b}{\cal O}_{ab}\ \alpha _a^{(n)}\ \alpha _b^{(n)}\ , \label{oss2} \end{equation} or in terms of the eigenvectors of the symmetric matrix $w$: \begin{equation} <{\cal O}(t)>=\sum_n\exp (\lambda _nt)\ \sum_{a,b}{\cal O}_{ab}\ (p_a^0\ p_b^0)^{1/2}\ e_a^{(n)}\ e_b^{(n)}\ . \label{oss3} \end{equation} In the following we report a detailed analysis of the equilibrium dynamics for a network of $400$ minima, averaging over $50$ different extractions of the parameters that define the model. We measure the time autocorrelation function of the stress tensor, the shear viscosity, the structural relaxation times and the mass diffusion coefficient. We first determine the time autocorrelation functions of a structural quantity which is well defined in all minimum configurations, i.e. the off-diagonal microscopic stress tensor (\ref{tsfo1}). The correlation function is \begin{equation} C(t)=\frac 13[<\sigma ^{zx}(t)\ \sigma ^{zx}(0)>+(xy)+(yz)]\ . \label{corre} \end{equation} The quantity ${\cal O}_{ab}$ in Eq. (\ref{me3}) is in this case \begin{equation} {\cal O}_{ab}=\frac 13[<\sigma _a^{zx}\ \sigma _b^{zx}>+(xy)+(yz)]\ . \end{equation} We have measured the correlation functions for different temperatures, from $T=150$ K to $T=20$ K. In Fig. $7$ we report the normalized correlation functions $C(t)/C(0)$ at different temperatures (open symbols) together with the best stretched exponential fit (lines): \begin{equation} C(t)=C(0) \; \exp \left[ -(t/\tau )^{\beta _k}\right] \ . \label{ste} \end{equation} Contrary to the MD computations, which result in a two-step behavior for the relaxation processes (one associated to fast local dynamics and the other to structural slow dynamics, the so called $\alpha $ structural processes), the model gives only one relaxation step, associated to the structural processes, because the model can only describes the long time behavior. The results we obtain with the present model are consistent with those of MD in the allowed region ( i.e. above $T\sim 90$ K, in order to avoid crystallization in the MD computation). In inset $(a)$ we show the temperature dependence of the stretching parameter $\beta _k$. It emerges that: the structural relaxation dynamics is well represented by a stretched exponential decay, and that the stretching parameter $\beta _k$ is strongly temperature dependent. Both results are well supported by experimental \cite{exp_beta} and numerical \cite{num_beta} observations. In our case the stretching parameters $\beta _k$ decreases from a value of $\sim 1$ at high temperature to $\beta _k\sim 0.35$ at low temperature, in agreement with experimental findings \cite{exp_beta}. From the behavior of the correlation function $C(t)$ we can obtain information about structural relaxation time $\tau $. The values of $\tau $ (inset $(b)$) are determined from the stretched exponential fits. We obtain an increase of many orders of magnitude in a small temperature range, as usually found in many glass-forming liquids. However, we do not find the dramatic increase of the Vogel-Tammann-Fulcher type expected for fragile glass-former \cite{ANGELL}. It is possible that the observed Arrhenius behavior emerges as a peculiar property of the model, which would mean that the model is unable to capture the phenomenology of ``fragility''. It is, however, possible that the Arrhenius law is a genuine property of glass-forming liquids with Lennard-Jones interaction, as supported by a comparison with a MD computation in the allowed temperature range. It would be very interesting to compare the behavior obtained from the model to the ``true'' (in the sense of MD computation) behavior in the full temperature range (this is possible only for those system that avoid crystallization, like suitable binary mixtures). From the time autocorrelation functions of the off-diagonal microscopic stress tensor, we can determine the shear viscosity as \cite{BALUCANI} \begin{equation} \eta =\frac 1{K_BTV}\int_0^\infty dt\ C(t)\ . \end{equation} Also in this case, like for the relaxation times, we find a strong increase in a small temperature range, from $\eta \sim 10^{-2}$ P at $T=150$ K, to $\eta \sim 10^{11}$ P at $T=20$ K. Also in this case we found a good agreement between the model and MD for $T>90$ K, giving further support to the model. The last quantity we measured in the model was the mass diffusion coefficient. In order to find it we determine the mean square displacement \begin{equation} {\cal O}(t)=\frac 1N|{\underline{r}}(t)-{\underline{r}}(0)|^2=\frac 1N \sum_{i=1}^N|{\vec r}_i(t)-{\vec r}_i(0)|^2\ , \end{equation} from which we obtain the diffusion coefficient: \begin{equation} D=\lim_{t\rightarrow \infty }\ \frac{{\cal O}(t)}{6t}\ . \end{equation} To evaluate $D$ we use the quantity ${\cal O}_{ab}=d_{ab}^2/N$, where $d_{ab}$ is defined in (\ref{dista}). We observe again a strong increase with an interesting behavior not simply linear in double logarithmic scale \cite{lettera}. We conclude this section by analyzing the validity of the Stokes-Einstein relation in the model. The Stokes-Einstein relation describes in a rigorous way the diffusive motion of a macroscopic objects in a fluids and predicts the following relation: \begin{equation} D\propto \frac T\eta \ . \label{se} \end{equation} The Stokes-Einstein relation also describes fairly well the diffusion at atomic scale in liquids at high temperatures. By lowering temperature, as observed in many experiments \cite{BSE}, one usually finds a breakdown of the Eq. (\ref{se}). We found \cite{lettera} that at high $T$ the model satisfies asymptotically the Stokes-Einstein relation, but upon decreasing $T$ we observe a breakdown of the relation and a fit over the lowest temperature data of the type \begin{equation} D^{-1}\propto (\frac \eta T)^\xi \ , \end{equation} gives the value \begin{equation} \xi \simeq 0.28\ . \end{equation} This value is in fairly good agreement with experimental results found in fragile glass-formers, like o-terphenyl \cite{FDSE}. \hfill \subsection{Off-equilibrium properties} Although introduced to analyze the long time dynamics, the model allows an easy computation of the short time, off-equilibrium dynamics. One of the main properties of glass-formers is the very strong increase of characteristic relaxation times when temperature is lowered. If these times become comparable to observational times, the system is no more able to explore the full accessible phase space and then to reach the thermal equilibrium. The observed quantities are characterized by off-equilibrium processes. In this regime the one-time quantities, such as energy or time correlation functions with fixed initial time, can no longer describe the physics of the system. The usual translational time invariance, valid in the equilibrium regime, is no more satisfied. One of the most interesting consequences of that is the fact that the fluctuation-dissipation relation does not longer hold\cite{BOUCH}. We concentrate here on this property. Let be $H$ the Hamiltonian of the system and ${\cal O}$ a generic observable dependent on microscopic variables. We define the two time autocorrelation function \begin{equation} C(t,t_w)=<{\cal O}(t)\ {\cal O}(t_w)>\ , \label{cor1} \end{equation} where we suppose $t>t_w$ and $<...>$ now means a dynamical average over initial conditions. We also introduce the response function to a perturbation $\epsilon $, which is coupled to the observable ${\cal O}$ and gives rise to a pertutbed Hamiltonian: \begin{equation} H^{\prime }=H+\epsilon (t)\ {\cal O}\ , \end{equation} The response is defined as \begin{equation} R(t,t_w)=\left. \frac{\delta <{\cal O}(t)>}{\delta \epsilon (t_w)}\right| _{\epsilon =0}\ , \end{equation} where again $t>t_w$. In the equilibrium regime the time translational invariance implies the validity of fluctuation dissipation relation \cite{PARSTA} \begin{equation} R_{eq}(\tau )=\beta \ \frac{\partial C_{eq}(\tau )}{\partial \tau }\ , \label{fde} \end{equation} where $\tau =t-t_w$. Introducing the integrate response function $\chi $ \begin{equation} \chi (t,t_w)=\int_{t_w}^tdt^{\prime }\ R(t,t^{\prime })\ , \end{equation} Eq. (\ref{fde}) takes the form \begin{equation} \frac{d\chi (C)}{dC}=\beta \ . \label{fdne3} \end{equation} In the off-equilibrium regime the fluctuation-dissipation ratio (\ref{fdne3}) is no more valid. It is possible however to generalize the ratio introducing a violation factor $X(t,t_w)$. The analytical study of some generalized mean field spin glass models \cite{CUGLI} shows that the function $X(t,t_w)$ depends on time only through the correlation function $C$: $X(t,t_w)=X[C(t,t_w)]$. Using this property we can write a generalized fluctuation-dissipation ratio in the off-equilibrium regime \begin{equation} \frac{d\chi (C)}{dC}=\beta \ X(C)\ . \label{fdne4} \end{equation} For short times $\tau \ll t_w$ we have $X(C)=1$ and the system satisfies an equilibrium-like relation, even if it is confined in a small phase space region. For times $\tau \sim t_w$ the exploration of the phase space is an off-equilibrium process and this implies the violation of the equilibrium fluctuation-dissipation ratio. In this case we have $X(C)<1$. The very interesting relationship between off-equilibrium and equilibrium properties of some generalized spin glass model, suggests that in the case of one step replica symmetry breaking the $X(C)$ function depends only on temperature \begin{equation} X(C)=m(T)\ , \end{equation} and $m(T)$ is linear in $T$ at low temperature. It has been recently suggested that structural glasses present a striking similarity with the generalized spin glass model with one step replica symmetry breaking \cite {equiv} (for a recent interesting review see Coluzzi \cite{COLUZZI}). We then expect that also for structural glasses the violation parameter would show a linear temperature dependence in the violation region $X<1$. Evidence of this behavior was found in recent numerical study of binary mixtures \cite {PABA}; we analyze it in our model. Let be ${\cal O}({\underline{r}}(t))$ a generic observable that depends on collective coordinates at time $t$. The average value of ${\cal O}$ in the off-equilibrium regime in the model is \begin{equation} <{\cal O}(t)>=\frac 1{M^{\prime }}\sum_{a,b}{\cal O}_a\ p_a(t;b,0)\ , \label{medne} \end{equation} where $a$ and $b$ label the minima and the sum over $b$ is now limited to a certain subset of minima. We chose the $M^{\prime }$ highest energy states. Expression (\ref{medne}) differs from the equilibrium one, as the initial states are weighted with a constant term (corresponding to an infinite temperature) rather than with the Gibbs-Boltzmann equilibrium weight. In this way we describe an instantaneous quench at time $t=0$ from $T=\infty $ to a finite temperature $T$ (the $T$ dependence is as usual in the probability $p_a$). The sum restricted to the $M^{\prime }$ initial states with highest energies ($M^{\prime }<M$ where $M$ is the total number of minima; in our case $M=400$ and $M^{\prime }=20$) allows a better description of the off-equilibrium regime. We calculate the time correlation functions in the model as \begin{equation} <{\cal O}(t)\ {\cal O}(t_w)>=\frac 1{M^{\prime }}\sum_{a,b,c}{\cal O}_a\ {\cal O}_b\ p_b(t_w;c,0)\ p_a(t;b,t_w)\ , \label{corne} \end{equation} where the sum over $b$ is still made over the $M^{\prime }$ highest minima. The quantity we determine is the time autocorrelation function of the off-diagonal microscopic stress tensor $\sigma ^{zx}$: \begin{equation} C(t,t_w)=<\sigma ^{zx}(t)\ \sigma ^{zx}(t_w)>-<\sigma ^{zx}(t)>\ <\sigma ^{zx}(t_w)>\ . \end{equation} The response function is determined by the perturbed Hamiltonian \begin{equation} H^{\prime }=H+\epsilon (t)\ \sigma ^{zx}\ , \end{equation} where the external field $\epsilon $ is \begin{equation} \epsilon (t)=\left\{ \begin{array}{ll} \displaystyle 0\displaystyle \ \ \ \ \ \ \ \ & \mbox{for \ \ $t<t_w$}\ , \\ & \\ \displaystyle \epsilon \displaystyle \ \ \ \ \ \ \ \ & \mbox{for \ \ $t\geq t_w$}\ . \end{array} \right. \label{pert} \end{equation} The perturbation induces a change in the energies of the minima: \begin{equation} \Phi _a^{^{\prime }}=\Phi _a+\epsilon (t)\ \sigma _a^{zx}\ . \end{equation} The response function is \begin{equation} R(t,t_w)=\left. \frac{\delta <\sigma ^{zx}(t)>_\epsilon }{\delta \epsilon (t_w)}\right| _{\epsilon =0}\ . \end{equation} The $<...>_\epsilon $ is evaluated in the presence of the perturbation $\epsilon $: \begin{equation} <\sigma ^{zx}(t)>_\epsilon =\frac 1{M^{\prime }}\sum_{a,b,c}\sigma _a^{zx}\ p_b(t_w;c,0)\ p_a^\epsilon (t;b,t_w)\ , \label{medper} \end{equation} where $p^\epsilon $ is the solution of the master equation with the perturbing term. For small perturbation we obtain: \begin{equation} \chi (t,t_w)=\frac{<\sigma ^{zx}(t))>_\epsilon -<\sigma ^{zx}(t))>_{\epsilon =0}}\epsilon \ . \end{equation} We have determined the correlation functions $C(t,t_w)$ and the response $\chi (t,t_w)$ as functions of $t$ for different times $t_w$; the temperatures we analyze are in the range $T=100\div 20$ K. In determining the response functions we have used a value of $\epsilon $ small enough ($\epsilon =0.1$) that the regime is linear, as verified by trying different $\epsilon $ values. In Fig. $8$ we report the behavior of $\chi $ versus $\beta C$ at temperatures $T=90$ K and $T=45$ K, respectively. While at $T=90$ K the relation between $\chi $ and $\beta C$ is to a good approximation linear with slope $1$ on the whole range (full line), at lower temperature it is evident that after a first linear behavior with slope $1$ (full line) an approximately linear behavior with slope $<1$ takes on at longer times (dashed line), as theoretically and numerically expected. Moreover the slope of the second region decreases by lowering temperature: in Fig. $9$ we show the slope $m$ of the violation region versus $T$. At high temperature the value of $m$ is nearly $1$, while below a temperature of about $60\div 70$ K $m$ decreases linearly, as we expect in the hypothesis of one step replica symmetry breaking. Fig. $9$ is limited to $T>40$ K, as for lower temperatures the $m$ values saturate to a limiting value and it is no more possible to extract correct information. This effect is probably due to the finite size of the system, because the sampling of the initial off-equilibrium states is not exhaustive ($M^{\prime }=20$). In the equilibrium analysis the finite size effects do not show up in the temperature range explored, as the sampling of the initial states is complete ($M^{\prime }=M$). In conclusion from the analysis of the off-equilibrium properties of the model it emerges that the deviation from the usual fluctuation-dissipation relation, valid in the equilibrium regime, is in agreement with theoretical predictions and numerical findings in simple glass-formers. \section{Conclusions} The very rich phenomenology of the cooling process of glass-forming liquids, of the glass transition and of the glassy systems in general, has received in the last few years many important theoretical, numerical and experimental contributions. The present work is concerned with the numerical investigation of a simple model glass, a Lennard-Jones system of interacting particles. The main aim of the work was to determine the emergent properties of the system at the level of potential energy landscape. After a detailed analysis of the topological properties of the potential energy surface, we introduced a model which reproduces the long time dynamic behavior of the system. While in the usual MD investigations of relaxation the computational times are proportional to physical times (with computational times of the order $10^5$ s one obtains physical times of the order $10^{-9}$ s for system of size $N\sim 10^3$), our model allows the study at very long physical times in short computational times. We studied both equilibrium and off-equilibrium properties. The main equilibrium results we obtained are ({\it i}) the stretching of the relaxation dynamics, ({\it ii}) the temperature dependence of the stretching parameter, and ({\it iii}) the breakdown of the Stokes-Einstein relation. If they are genuine properties of the glassy system analyzed, they represent intriguing and interesting results that open fascinating questions about the glassy and supercooled liquid behavior. Although introduced to investigate the long time dynamics, the model is also able to describe in a simple and direct way the off equilibrium dynamics. The emergent violation of the fluctuation-dissipation relation (that holds at equilibrium) is a very interesting feature and supports many conjectures about the analogy between structural glasses and some spin glass model \cite{equiv}. Moreover, the appearance of a critical temperature below which the violation takes place, seems to indicate the existence of a transition; its relation with the equilibrium properties is an open and exciting question. In conclusion, the analyzed features of the potential energy landscape and the emergent properties of the model both at and off equilibrium, seem to provide a good description of the glassy systems. The method is very powerful for the investigation of the glassy properties by avoiding some of the main problems usually encountered in numerical studies, like the very long computational times in the low temperature regime or the presence of crystal states. We hope the analysis we performed may constitute a promising route in the investigation of glassy systems. \hfill We acknowledge B.~Coluzzi, G.~Monaco, F.~Sciortino and P.~Verrocchio for useful discussions, and D.~Leporini who kept our attention on the fractional SE relation issue.
\section{Introduction} The isolated nonthermal filaments (hereafter, NTFs) in the Galactic center (hereafter, GC) have remained unexplained since their discovery by Morris and Yusef-Zadeh (1985). It is generally accepted that these are magnetic structures emitting synchrotron radiation since their emission is strongly linearly polarized with the magnetic field generally aligned with the long axis of the filaments (Bally \& Yusef-Zadeh 1989; Gray et al. 1995; Yusef-Zadeh, Wardle, \& Parastaran 1997; Lang, Morris, \& Echevarria 1999). These structures are notable for their exceptionally large length to width ratios, of order 10 to 100, and remarkable linearity (Yusef-Zadeh 1989; Morris 1996). To date, seven objects have been classified as NTFs. Six of these point perpendicularly to the Galactic plane, but the most recently discovered NTF is parallel to the plane (Anantharamaiah et al 1999). The filaments have lengths up to 60 pc and often show feathering and sub-filamentation on smaller transverse scale when observed at high spatial resolution (Liszt and Spiker 1995; Yusef-Zadeh, Wardle, \& Parastaran 1997; Lang, Morris \& Echevarria 1999; Anantharamaiah et al 1999). The observed radio 20/90 cm spectral indices (defined as the source flux, $S$, varying as $\nu^{\alpha}$) show a range, $-0.3 < \alpha < -0.6$ (LaRosa et al 1999). To date, there is no strong evidence that the spectral index varies as a function of length along the NTF (Lang, Morris \& Echevarria 1999; Kassim et al. 1999). Lastly, it appears that all well studied NTFs may be associated with molecular clouds and/or H II regions (Serabyn \& Morris 1994; Uchida et al. 1996; Stahgun et al 1998). Several different types of models have been proposed for the filaments. These include magnetic field generation by an accretion disk dynamo with subsequent transport of field into the interstellar medium (Heyvaerts, Norman, \& Pudritz 1988); electrodynamic models of molecular clouds moving with velocity ${\bf v}$ across a large scale ordered magnetic field, ${\bf B}$, resulting in current formation by ${\bf v \times B}$ electric fields and subsequent pinching of these currents into filaments (Benford 1988, 1997; Lesch \& Reich 1992); magnetic reconnection between a molecular cloud field and the large-scale ordered field (Serabyn \& Morris 1994); and particle injection into interstellar magnetic field ropes at a stellar wind termination shock (Rosner and Bodo 1996). Nicholls and Le Stange (1995) proposed a specifically tailored model for G359.1-0.2, also called the ``Snake''. They invoke a high velocity star with a strong stellar wind that is falling through the galactic disk, to create a long wake, which they call a ``star trail''. They must, however, fine tune their model in order to obtain radio emission from the trail by requiring the high energy electrons to be injected into the trail from the supernova remnant G359.1-0.5. Although each of these models can in principle explain the particle acceleration and radio emission, none except for the specialized star trail scenario has satisfactorily accounted for the observed structure of the filaments. For instance, Rosner and Bodo (1996) employ a stellar wind termination shock as the source of high energy particles that they assume are loaded onto pre-existing interstellar field lines. The width of the resulting NTF is the radius of the stellar wind bubble. Synchrotron cooling leads, through a thermal instability, to collapse of the filaments and amplification of the internal magnetic field. The streaming of the particles along these otherwise quiescent field lines is assumed to produce the observed long threads. However, MHD stability is a problem for this model, and indeed for all models listed above, because magnetic fields left to their own devices will deform through a rich variety of modes. These range from kink and sausage instabilities for ideal MHD to tearing modes for resistive plasmas (e.g. Parker 1977; Cravens 1997). Unless very special magnetic field configurations and boundary conditions are imposed, and these are difficult to achieve even in the laboratory, the length and thinness of the NTFs cannot be explained as static structures. In this paper, we adopt the viewpoint that the filaments are not equilibrium structures but are rather dynamical structures embedded in a flow. In a flow, the growth of many of the local instabilities is suppressed by the advection. A similar conclusion has been reached by Chandran, Cowley, \& Morris (1999) who argue that the filaments represent locally illuminated regions of a large scale strong magnetic field that is formed through the amplification of a weak {\it halo} field by a galactic accretion flow. In contrast, we propose an alternative model in which the advection of a weak {\it galactic} field in a large scale outflow from the central region is amplified locally by encounters with interstellar clouds. We find that several key elements of the previously published scenarios are the natural consequences of this cloud-wind interaction picture. \section{The Comet Model} The key feature of the physical interaction between a comet and the magnetized solar wind was identified by Alfv\'en (1957) and elaborated by many subsequent studies (e.g. Russell et al. 1991; Luhmann 1995; Cravens 1997). As the magnetic field that is carried in the wind plasma encounters the comet, the field progress is retarded through the coma because the magnetic diffusion times are much longer than the advective timescale. Mass loading of the solar wind from the coma produces a velocity gradient, $\partial v_w/\partial x$, where $x$ is the cross-tail direction. The external field drapes over the coma and is stretched by the wind, ultimately forming a current sheet in the antisolar direction. This field line draping, for a molecular cloud, is depicted in Figure 1. Remote observations show that cometary streamers routinely display aspect ratios of a 100 or more (Jockers 1991). Direct {\it in situ} plasma measurements of comet Giacobini-Zinner have confirmed the overall picture of magnetic field draping. In particular, these encounter measurements show that the central tail axis consists of a plasma sheet with very low magnetic field (Siscoe et al. 1986). This sheet is surrounded by a low density plasma that is threaded with the draped wind magnetic field that has been compressed and amplified by the flow. Transverse pressure balance requires that the draped field is about a factor of 5 to 10 stronger than the ambient field with amplification occurring because of flux conservation and field line stretching. This is our basic cometary analogy, that the ambient field is anchored in the cloud and that the field line tilt and amplification result from the shearing between the wake and the external wind. This picture, which explains solar system scale phenomena rather well, is more than a mere analogy. Any magnetized wind that impacts a finite blunt body with low resistivity will deflect around the object and drape the field along the wake flow. The field diffusion time is $t_d = L^2/\eta$, where here $L$ is the cloud radius and $\eta$ is the resistivity. For the GC, this timescale is many orders of magnitude larger than the wind advection time. Consequently, the field evolution around the molecular cloud is similar to the cometary case provided the field is ordered on the cloud size scale, $L$. We now explore the consequences of this scenario for the NTFs. Consider a galactic scale wind with a mass loss rate $\dot{M}$. This wind need not eminate only from the GC. With the broad spatial distribution of star forming regions in the inner galaxy, we would anticipate a roughly cylindrical -- not radial -- outflow which would be sampled by clouds moving in whatever orbits they happen to have relative to the plane. Consequently, the wakes so produced should generally be perpendicular to the plane. For simplicity, however, we will assume here a compact source. The number density in the wind is given by $n_W = \dot{M}v_{w,3}^{-1}r_{100}^{-2}$ cm$^{-3}$, for a mass loss rate in $M_\odot$yr$^{-1}$, a wind speed $v_{w,3}$ in $10^3$~km$\;$s$^{-1}$, and a distance $r_{100}$ in 100 pc. For a cloud to survive in a postulated galactic scale wind, its internal pressure must at least balance the ram pressure of the background. We assume that the cloud pressure is given by $P= \rho_c\sigma_c^2$ where $\rho_c$ is the cloud mass density and $\sigma_c$ its internal velocity dispersion. Hence, for a wind of density $\rho_w$ and speed $v_w$, the required cloud density is given by $\rho_c = \rho_w(\sigma/v_w)^2$. It has been inferred from {\it Ginga} and ASCA X-ray observations that the inner Galaxy displays a strong wind (Yamauchi et al. 1990; Koyama et al. 1996). The average wind density within a radius of 80 pc of the GC is around 0.3 cm$^{-3}$ with a temperature of 10 keV and an expansion velocity of about 3000 ~km$\;$s$^{-1}$ (Koyama et al. 1996). These parameters correspond to a mass loss rate of 10$^{-2}$ M$_\odot$ yr$^{-1}$ for a wind speed of around 1000 km s$^{-1}$, which yields a critical cloud density of order $n_c \ge 10^3$ cm$^{-3}$ for $\sigma = 20$ ~km$\;$s$^{-1}$ (a typical linewidth for molecular clouds in the GC region, see Morris \& Serabyn [1996]) although clouds nearer the center will need higher densities to survive. This density estimate is a lower limit. For clouds to survive in the GC tidal field they must have densities at least an order of magnitude above this (e.g. G\"usten 1989). The effect on the cloud population is that massive, dense clouds will survive while lower density, low mass clouds likely disperse on a dynamical timescale and thus the cloud population may depend on galactocentric distance. Dense clouds form wakes by geometrically blocking and deflecting the wind. We identify this wake, drawn out by the wind, with the NTFs. This scenario is sketched in Figure 1. Thus follows the essential predictive feature of our model: since the filaments are not static structures, the classic MHD instabilities do not limit the aspect ratio as they would for a static equilibrium field. What determines the structural properties of the wake, {\it i.e.} its aspect ratio and length? Given that $t_d \gg L/v_w$, the draped field is stretched by the wind. If $\delta v$ is the boundary layer shear between the wind and cloud wake and $\Delta$ is a characteristic length for the layer (of order $L$), then the axial field, $B_z$, as a function of distance $Z$ behind the cloud is given by the induction equation, $\partial B_z/\partial t = (\partial v/\partial x)B_0$, where $B_0$ is the external field. This has the approximate solution: \begin{equation} B_z = B_0\frac{\delta v}{\Delta}\frac{z}{v_w}. \end{equation} The axial field will continue to amplify until the draped magnetic field pressure balances the ram pressure of the wind. In other words, when the wake Alfv\'en speed equals the wind speed, $B_z/(4\pi \rho_0)^{1/2} = v_w$, the field can no longer be stretched. This provides a critical length: \begin{equation} z_c = 5\times 10^2n^{1/2}v_{w,3}^2\Delta /(B_{0,\mu}\delta v_3), \end{equation} where $v_{w,3} = v_w/10^3$km s$^{-1}$, $n$ is the number density, and $B_{0,\mu}$ is the external field in $\mu$G. Thus, for $n \approx 1$ cm$^{-3}$ and $B_{0,\mu} \approx 10$, the predicted aspect ratio is $z_c/\Delta \approx 50$. Notice that stronger ambient fields lead to shorter wakes. We now address the question of stability for the filaments. In the MHD case, the velocity shear must exceed the Alfv\'en speed to produce a growing mode for the Kelvin-Helmholtz instability (KHI). Other classical instabilities, such as the streaming, sausage, and kink modes, have a similar criterion (e.g. Wang 1991). Nonlinear models by Malagoli, Bodo \& Rosner (1996) find that the fastest growing mode has a wavenumber given by $k\Delta \approx 0.05$. The KHI can therefore be suppressed if the draped field amplification length is less than $2\pi/k$. Hence, for stability $z_c \le 40\pi \Delta$. With this constraint, we find a lower limit on the external wind field strength, $B_{0,\mu} \ge 40 n^{1/2}v_{w,3}/\delta v_3$. Equipartition for the wind plasma gives $B_{0,\mu} \approx 20$ for the parameters derived from the ASCA data, which is in surprisingly good agreement with the stability constraint. Thus the expected amplified field strength is $B_z \approx 2$ mG for $z/\Delta$ given by eq. (2). Thus the key parameters that can be derived from our model are the aspect ratio which depends on the wind parameters, and the magnetic field strength in the filament. The observed aspect ratios can be explained using equation (2) with wind parameters consistent with the ASCA data. There are no direct measurments of the magnetic fields in the filaments. An estimate for the magnetic field can be derived from the observed synchrotron luminosities using a minimum energy analysis. The synchrotron luminosities are around $10^{33} - 10^{34}$ erg s$^{-1}$ (Gray et al. 1995; Lang, Morris, \& Echevarria 1999; Kassim et al. 1999) and yield a magnetic field of $\sim 0.1$ mG about an order of magnitude smaller than our model result. Another estimate for the field strength comes from assuming that the particles traverse the length of an NTF in a time equal to their synchrotron lifetime. The synchrotron lifetime is $t_{\frac{1}{2}} = 1.20\times 10^4 B_{z,mG}^{-2}E_{GeV}^{-1}$ yrs, where $B_{z,mG}$ is the axial field in milligauss and $E$ is the electron energy in GeV (e.g. Moffatt 1975). Without reacceleration, assuming that the electrons are injected near one end and radiate as they stream at the Alfv\'en speed (e.g. Wentzel 1974), the observed filament lengths give a field strength of 1 mG for a length scale of 30 pc. Fields strengths of 1 mG have also been derived from dynamical arguments by Yusef-Zadeh and Morris (1987). We therefore conclude that our estimate of 1 mG is very reasonable and that the minimum energy analysis of such structures which assumes static and/or equilibrium conditions may produce misleading results. Note that the synchrotron lifetime arguement indicates that reacceleration or acceleration along the length of the filament is not required, although as we now discuss acceleration along the filament is expected in our picture. Finally, since the NTFs are radiating via synchrotron emission, we address the question of particle energization. The observed emission requires only a very small population of relativistic particles, of order $10^{-5}$cm$^{-3}$. The maximum energy that is available for conversion to high energy particles is $VB_z^2/8\pi$, where $V$ is the volume of the wake. The maximum mean energy per particle that results from this conversion is 10 GeV. This is more than enough to explain the radio emission. A number of mechanisms that may be responsible for particle acceleration are natural consequences of this MHD configuration. The wake must contain a current sheet. Such structures have been extensively studied in space plasmas. The simulation of sheared helmet plumes in the solar corona by Einaudi et al. (1999) is particularly relevant to our scenario. They show that a current sheet imbedded in a wake flow is unstable to the generation of a local turbulent cascade without destruction of the large scale advected structure. Such cascades efficiently accelerate particles through wave-particle interactions (Miller et al. 1997). This turbulent acceleration would therefore occur along the entire length of the filament, thus spectral aging would not be observed in this scenario. \section{Discussion and Conclusions} Santillan et al. (1999) have recently published numerical MHD simulations of cloud collisions with a magnetized galactic disk. Although these are ideal MHD and not of wind flow, they clearly demonstrate that field line draping occurs as the interstellar clouds move through a background large scale field. In particular, their Fig. 4 shows the formation of a narrow straight tail for the cloud slamming into a transverse field imbedded in a planar gas layer. Dynamically, this simulation differs from wind flow because the cloud is slowed by the environmental gas. Yet the essential physical process is the same and closely resembles the simulations of cometary tail evolution by Rauer et al. (1995). This cloud-wind interaction, which may destroy the clouds if their masses are low enough (see Vietri, Ferrara, \& Miniati 1997), is able to generate long magnetized tails with large aspect ratios. Nonetheless, it has been argued in the literature that there is no need for a dynamical explanation of these structures. Recalling our discussion of the various proposed static models for the NTFs, the common explanation for their stability hinges on the existence of a pervasive background field. We see two ways of interpreting the magnetic field measurements obtained from the filaments. One is to assume that they represent local enhacements of an otherwise weak, but invisible, field. The other is to assume that one is seeing a region that happens to be locally illuminated but is otherwise extensive and uniform. We explicitly adopt the local enhancement picture and propose a dynamical mechanism that can amplify the field to much higher strength and still be stable. On the other hand, assuming a pervasive field still leaves the stability question unresolved for the following reasons. A force-free equilibrium background field that is presumably anchored in the turbulent gas of the galactic center certainly will not be stable. For instance, the solar corona has a pervasive field that suffers both local and global instabilites. Moreover, to stablize a filament, a pervasive field must have a pressure gradient perpendicular to the filaments, so the field cannot be uniform. If it has gradients and curvature, a static magnetic field is likely to be unstable. In contrast, stability is not an issue for a dynamical model, whether the flow is accreting (Chandran et al. 1999) or, as in our case, an outflowing wind. The simplest geometry predicted by the cometary analogy is that every filament should be associated with a molecular cloud on the side toward the galactic plane. This is seen for the Sgr C filament (Liszt \& Spiker 1995) and the ``Snake'' (Uchida et al. 1996). The model does not, however, require this and more complex geometrical arrangements are certainly possible in which environmental clouds interact with or are superimposed on the filaments almost anywhere along their lengths. For instance, Yusef-Zadeh \& Morris (1987) find that a milligauss field suffices to stabilize the filaments against ram pressure by colliding molecular clouds. We note that this is precisely the field strength produced dynamically by the cometary model. In addition, a final state, where the cloud is completely dissipated, could still permit the survival of the filament and has a cometary analog. There are many instances in comets where the tail completely separates from the coma and yet maintains structural coherence as it is advected in the solar wind (e.g. Brandt \& Niedner 1987). These so-called disconnection events could also occur in our picture. In such instances, there would be no cloud at either end of the filament. We close by emphasizing that our aim here has been the exploration of the consequences of a general scenario that can serve as a framework for more quantitative calculations of the physical properties of the Galactic Center filaments. Although we use the special conditions at the GC to constrain the mechanisms, the model is not constructed specifically to explain the NTFs. Instead, they result from the conditions that likely arise in any starburst galactic nucleus (see Mezger, Duschl, \& Zykla 1996) and should be observable in such environments. \acknowledgments We thank G. Einaudi, J. R. Jokipii, N. Kassim, C. Lang, J. Lazio, M. Niedner, and M. Vietri for discussions, and A. Santillan for permission to quote his results prior to publication. We especially thank the referee, Mark Morris, for his critical reading of the manuscript and for discussions, and B. D. G. Chandran for communicating his paper in advance of publication. TNL is supported by a NAVY-ASEE faculty fellowship from the Naval Research Laboratory and a NASA JOVE grant to Kennesaw State University. SNS is partially funded by NASA and thanks the Astrophysics Group of the Physics Department of the University of Pisa for a visiting appointment during summer 1998.
\section{INTRODUCTION} \label{sec:Introd} The stage leading up to dynamic collapse of a magnetically subcritical cloud core to a protostar or a group of protostars is believed to be largely quasi-static, if the responsible process is ambipolar diffusion (e.g., Mestel \& Spitzer 1956, \cite{Nakano79}, \cite{LS89}, \cite{Tomisaka89}, \cite{BM94}).\footnote{For example, as measured either by the net accelerations or by the square of the inward flow speed divided by the sound speed, the ambipolar-diffusion models in Figures 3 and 6 of \cite{CM94} spend less than 0.1\% of the total computed evolutionary time in states where even a single grid point is more than 10\% out of mechanical balance with self-gravity, magnetic forces, and thermal pressure (see also Figs. 3 and 7 of Basu \& Mouschovias~1994 and Fig.~1 of Ciolek \& Koenigl~1998).} To describe the transition between quasi-static evolution by ambipolar diffusion and dynamical evolution by gravitational collapse, \cite{LS96} introduced the idea of a pivotal state, with the scale-free, magnetostatic, density distribution approaching $\rho \propto r^{-2}$ for an isothermal equation of state (EOS) when the mass-to-flux ratio has a spatially constant value, a condition that \cite{SL97} and \cite{LS97} termed ``isopedic''. Numerical simulations of the contraction of magnetized clouds justify the assumption of a nearly constant mass-to-flux ratio in the pivotal core.\footnote{For example, inside the starred point where \cite{CM94} consider the core to begin, the mass-to-flux ratio varies in the last models of their Figures 3 and 6 by a factor of only 3 or 2 over a range where the density varies by a factor $\sim 10^5$. Outside the starred point, the mass-to-flux value exhibits greater variation, but this occurs only because \cite{CM94} impose starting values for the mass-to-flux in the envelope that are $\sim 2\times 10^{-2}$ times the critical value (see also Figs. 4a and 8b in \cite{BM94}). Such small ratios for the bulk of the mass of a molecular cloud are probably ruled out by the Zeeman OH measurements summarized by Crutcher (1998).} The small dense cores of molecular clouds that give rise to low-mass star formation are effectively isothermal (\cite{MB83}; \cite{SAL87}). The situation may be different for larger regions that yield high-mass or clustered star formation. It has often been suggested that the EOS relating the gas pressure $P$ to the mass density $\rho$ of interstellar clouds can be represented by a polytropic relation $P \propto \rho^{1 + 1/n}$ with negative index $n$. \cite{Shu72} pointed out the utility of this idealization within the context of the classic two-phase model of the diffuse interstellar medium [\cite{Pikelner67}; \cite{FGH69}; \cite{SpitzerandScott69}], while \cite{VialaandHoredt74} published extensive tables analyzing the stability of non-magnetized, self-gravitating spheres of such gases. \cite{Maloney88} examined the linewidth-size and density-size relations of molecular clouds, first found by \cite{Larson81} and subsequently studied by many authors [e.g., \cite{Leung82}, \cite{Torrelles83}, \cite{Dame86}, \cite{Falgarone92}, \cite{MieschB94}]. Maloney pronounced the results consistent with the properties of negative index polytropes. For a polytropic EOS, the sound speed $c_s\equiv (dP/d\rho)^{1/2}\propto \rho^{1/2n}$ increases with decreasing density if $n<0$. The latter behaviour may be compared with the empirical linewidth-density relation for molecular clouds, $\Delta v\propto \rho^{-q}$, with $q \simeq 0.5$ for low-mass cores (\cite{MF92}) and $q \simeq 0.15$ for high-mass cores (\cite{CaM95}), implying that $n$ lies between $-1$ and $-3$, or that a static $\Gamma \equiv 1+1/n$ lies between $0$ and 0.7. The case $\Gamma=1/2$ is of particular relevance for the equilibrium properties of molecular clouds. \cite{Wa44} found that the pressure of Alfv\`en waves propagating in a stratified medium, $P_{\rm wave}\propto |\delta{\bf B}|^2$, in the absence of damping obeys the simple polytropic relation $P_{\rm wave} \propto \rho^{1/2}$, a consequence of conservation of the wave energy flux $v_A |\delta{\bf B}|^2$. This result was later derived more rigorously by \cite{W62} in the WKB approximation for MHD waves propagating in mildly inhomogeneous media, and, more recently by \cite{FA93} and \cite{McKZ95} in a specific astrophysical context. In numerical simulations of the same problem, \cite{GO96} found indication of a much shallower relation ($\Gamma \simeq 0.1$) for a self-gravitating medium supported by nonlinear Alfv\`en waves. On the other hand, for the adiabatic contraction of a cloud supported by linear Alfv\`en waves, \cite{McKZ95} found a dynamic $\gamma$ larger than 1. \cite{Vazquez97} confirmed a similar behaviour in numerical simulations of the gravitational collapse of clouds with an initial field of hydrodynamic rather than hydromagnetic turbulence. In the limit of $\Gamma\rightarrow 0$ (or $n\rightarrow -1$), the EOS becomes ``logatropic,'' $P\propto \ln\rho$, a form first used by \cite{LS89} to mimic the nonthermal support in molecular clouds associated with the observed supersonic linewidths. The sound speed associated with the nonthermal contribution, $c_s = (d P / d \rho)^{1/2} \propto \rho^{-1/2}$ becomes important at the low densities characteristic of molecular cloud envelopes (as contrasted with the cloud cores) since the thermal contribution is independent of density if the temperature $T$ remains constant. This nonthermal contribution decreases with increasing density and will become subsonic at high densities as recently observed in the central regions of dense cores (\cite{BG98}). \cite{McLP96} and \cite{McLP97} have modeled the equilibrium and collapse properties of unmagnetized, self-gravitating, spheres with a pure logatropic EOS and claim to find good agreement with observations. Adams, Lizano, \& Shu (in 1987) independently obtained the similarity solution for the gravitational collapse of an unmagnetized singular logatropic sphere (SLS), but they chose not to publish their findings until they had learned how to magnetize the configuration in a nontrivial way (see the reference to this work in Fuller \& Myers~(1992), who considered the practical modifications to the protostellar mass-infall rate introduced by ``nonthermal'' contributions to the support against self-gravity). Magnetization constitutes an important program to carry out if we try to justify a nonthermal EOS as the result of a superposition of propagating MHD waves (see also Holliman \& McKee 1993). In this paper, we extend the study of Li \& Shu (1996) to include the isopedic magnetization of pivotally self-gravitating clouds with a polytropic equation of state. As a by-product of this investigation, we obtain the unanticipated and ironic result that the only way to magnetize a singular logatropic configuration and maintain a scale-free equilibrium is to do it trivially, i.e., by threading the SLS with straight and uniform field lines (see \S 6). A basic consequence of treating the turbulence as a scalar pressure, coequal to the thermal pressure except for satisfying a different EOS, is that we do not change the basic topology of the magnetic field. This assumption may require reassessment if MHD turbulence enables fast magnetic reconnection (\cite{VL99}) and allows the magnetic fields of highly flattened cloud cores (Mestel \& Strittmatter 1967, \cite{BM96}) or pseudodisks (\cite{GS93b}) to disconnect from their background. Recent MHD simulations carried out in multiple spatial dimensions (e.g., \cite{sto98}; \cite{mac98}; \cite{ost99}; \cite{pad99}) find turbulence in strongly magnetized media to decay almost as fast as in unmagnetized media. Such decay may be responsible for accelerating molecular cloud core formation above simple ambipolar diffusion rates (\cite{Nakano98}, \cite{ML98}, \cite{shuetal99}). Although this result also cautions against treating turbulence on an equal footing as thermal pressure, we attempt a simplified first analysis that includes magnetization to assess the resulting configurational changes when we adopt an alternative EOS for the pivotal state. In particular, different power-law dependences of the radial density profile translate immediately to different time dependences in the mass-infall rate for the subsequent inside-out collapse (\cite{Cheng78}, McLaughlin \& Pudritz~1997). The paper is organized as follows. In \S 2 we formulate the equations of the scale-free problem and show that each solution depends only on the polytropic exponent $\Gamma$ and a nondimensional parameter $H_0$ related to the cloud's morphology. In \S 3 we present the numerical results. In \S 4, \S 5, and \S 6 we discuss the limiting form of the solutions. Finally, in \S 7 we give our conclusions and discuss the possible implications of our results for star formation and the structure of giant molecular clouds. \section{SELF-SIMILAR MAGNETOSTATIC EQUILIBRIUM EQUATIONS} To begin, we generalize the singular polytropic sphere in the same way that \cite{LS96} generalized the singular isothermal sphere (SIS). In the absence of an external boundary pressure, the only place the pressure $P$ enters in the equations of magnetostatic equilibrium is through a gradient. Consider then the polytropic relation \begin{equation} \label{dpdrho} {dP\over d\rho}=K\rho^{-(1-\Gamma)}. \end{equation} By integrating equation (\ref{dpdrho}) we recover for $\Gamma = 1$ the isothermal EOS, $P= K \rho$, (where $K$ is the square of the isothermal sound speed) and for $\Gamma = 0$ the logatropic EOS, $P = K \ln\rho$. We adopt axial symmetry in spherical coordinates and consider a poloidal magnetic field given by \begin{equation} \label{bfield} {\bf B} = {1 \over 2 \pi} \nabla \times \left( {\Phi \over r \sin \theta} \hat e_\phi \right), \end{equation} where $\Phi(r,\theta)$ is the magnetic flux. Force balance along field lines requires \begin{equation} V+{1 \over \Gamma - 1} K\rho^{-(1 - \Gamma) }=h(\Phi), \end{equation} where $V$ is the gravitational potential and $h(\Phi)$ is the Bernoulli ``constant'' along the field line $\Phi =$ constant. Poisson's equation now reads \begin{equation} \label{along} {1\over r^2}{\partial\over\partial r}\left[r^2\left({dh\over d\Phi} {\partial\Phi\over\partial r}- K\rho^{-(2-\Gamma)}{\partial\rho\over\partial r} \right)\right]+{1\over r^2\sin^2\theta} {\partial\over\partial\theta}\left[\sin\theta\left({dh\over d\Phi} {\partial\Phi\over\partial\theta}-K\rho^{-(2-\Gamma)} {\partial\rho\over\partial\theta}\right)\right]=4\pi G\rho; \end{equation} whereas force balance across field lines reads \begin{equation} \label{across} {1\over 16\pi^3r^2\sin^2\theta}\left({\partial^2\Phi\over\partial r^2}+ {1\over r^2}{\partial^2\Phi\over\partial\theta^2}-{\cot\theta\over r^2} {\partial\Phi\over\partial\theta}\right)=-\rho{dh\over d\Phi}. \end{equation} We look for scale-free solutions of the above equations by nondimensionalizing and separating variables: \begin{mathletters} \begin{equation} \rho=\left({K\over 2\pi Gr^2}\right)^{1/(2-\Gamma)}R(\theta), \end{equation} \begin{equation} \Phi=4\left({\pi^{3-2\Gamma}Kr^{4-3\Gamma}\over G^{\Gamma/2}} \right)^{1/(2-\Gamma)}\phi(\theta), \end{equation} \begin{equation} \label{nondim} {dh\over d\Phi}=H_0\left(2^{3\Gamma-2}KG^{2-2\Gamma}\over \pi^{1-\Gamma} \Phi^{2-\Gamma}\right)^{1/(4-3\Gamma)}, \end{equation} \end{mathletters} where $H_0$ is a dimensionless constant that measures the deviation from a force free magnetic field, and $R(\theta)$ and $\phi(\theta)$ are dimensionless functions of the polar angle $\theta$.\footnote{These definitions are not applicable for $\Gamma = 4/3$ or $\Gamma=2$.} These assumptions imply that the equilibria will have spatially constant mass-to-flux ratios (see below). Substitution of equation (\ref{nondim}) into equations (\ref{along}) and (\ref{across}) yields $$ {1\over\sin\theta}{d\over d\theta}\left\{\sin\theta \left[A_\Gamma H_0\phi^{-(2-\Gamma)/(4- 3\Gamma)}\phi^\prime- R^{-(2-\Gamma)}R^\prime\right]\right\}= $$ \begin{equation} \label{alongn} 2\left[R-{(4-3\Gamma) \over(2-\Gamma)^2} R^{-(1-\Gamma)}-\left({4 - 3\Gamma\over 2 - \Gamma}\right)^2B_\Gamma H_0 \phi^{2(1-\Gamma)/(4-3\Gamma)}\right], \end{equation} and \begin{equation} \label{acrossn} {1\over\sin^2\theta}\left[\phi^{\prime\prime}-\cot\theta\phi^\prime+ {2(4-3\Gamma)(1-\Gamma)\over(2-\Gamma)^2}\phi\right]=-C_\Gamma H_0R \phi^{-(2-\Gamma)/(4-3\Gamma)}, \end{equation} where a prime denotes differentiation with respect to $\theta$, and \begin{mathletters} \begin{equation} A_\Gamma=2^{\Gamma(3-2\Gamma)/(4-3\Gamma)(2-\Gamma)}, \end{equation} \begin{equation} B_\Gamma=2^{-(1-\Gamma)(8-5\Gamma)/(4-3\Gamma)(2-\Gamma)}, \end{equation} \begin{equation} C_\Gamma=2^{-\Gamma(1-\Gamma)/(4-3\Gamma)(2-\Gamma)}. \end{equation} \end{mathletters} In particular, for $H_0=0$, eq.~(\ref{alongn}) gives the dimensionless density for the non-magnetized singular polytropic sphere \begin{equation} \label{nonmagR} R = \left[ {4-3\Gamma\over (2-\Gamma)^2}\right]^{1/(2-\Gamma)}, \end{equation} whereas eq.~(\ref{acrossn}) implies $\Phi=0$ for $0 < \Gamma \le 1$, in order to satisfy the boundary conditions eq.~(\ref{bc}). In this case, the mass-to-flux ratio $\lambda_r$ is infinite. However, for $\Gamma=0$, eq.~(\ref{acrossn}) admits also the analytic solution of $\Phi \propto r^2 \sin^2\theta$ corresponding to a straight and uniform field, while the density function is $R(\theta) = 1$. Therefore, a spherical logatropic scale free cloud can be magnetized with a uniform magnetic field of any strength, and any value of the spherical mass to flux ratio is allowed.\footnote{In this case, $\lambda_r^2=2\mu^2=[2\phi(\pi/2)^2]^{-1}$.} For arbitrary values of $\Gamma$ and $H_0$ the ordinary differential equations (ODEs) (\ref{alongn}) and (\ref{acrossn}) are to be integrated subject to the two-point boundary conditions (BCs): $$ \lim_{\theta\rightarrow 0}\sin\theta\left[A_\Gamma H_0\phi^{-(2-\Gamma)/(4-3\Gamma)} \phi^\prime-R^{-(2-\Gamma)} R^\prime\right]=0, $$ \begin{equation} \label{bc} \phi(0)=0,~~~\phi^\prime(\pi/2)=0,~~~R^\prime(\pi/2)=0. \end{equation} The first BC implies that there is no contribution from the polar axis to the mass inside a radius $r$. The second BC comes from the definition of magnetic flux, i.e. no trapped flux at the polar axis. The last two BCs imply no kinks at the midplane. The equilibria are characterized by: \noindent ({\em a}\/) the spherical mass-to-flux ratio, \footnote{The standard mass-to-flux ratio $\lambda = 2 \pi G^{1/2} M(\Phi)/ \Phi$ is not defined for the polytropic scale free magnetized equilibria because the integral $\int_0^{\pi/2} R(\theta) \phi(\theta)^{-1}\sin\theta d\theta$ diverges since it can be shown that $R(\theta=0) \ne 0$ for $\Gamma <1$.} \begin{equation} \label{lambdar} \lambda_r\equiv 2 \pi G^{1/2} { M(r) \over \Phi(r,\pi/2)} = 2^{(1-\Gamma)/(2-\Gamma)} \left( {2 -\Gamma \over 4 - 3\Gamma} \right ) {1 \over \phi(\pi/2)} \int_0^{\pi/2}R(\theta)\sin\theta d\theta , \end{equation} where $M(r)$ is the mass enclosed within a radius $r$; \noindent ({\em b}\/) the factor $D$ by which the average density is enhanced over the non-magnetized value because of the extra support provided by magnetic fields, \begin{equation} \label{D} D \equiv \left[ {4-3\Gamma\over (2-\Gamma)^2}\right]^{-1/(2-\Gamma)} \int_0^{\pi/2}R(\theta)\sin \theta \, d\theta , \end{equation} which is equal to $1$ if $H_0=0$ (see eq.~[\ref{nonmagR}]); \noindent ({\em c}\/) the sound speed, \begin{equation} \label{cs} c_s^2 = \left( 2 \pi G r^2 \right)^{(1-\Gamma)/(2-\Gamma)} K^{1/(2-\Gamma)} R(\theta)^{-(1-\Gamma)}; \end{equation} and ({\em d}\/) the Alfv\`en speed \begin{equation} v_A^2=2^{\Gamma/(2-\Gamma)} \left( 2 \pi G r^2 \right)^{(1-\Gamma)/(2-\Gamma)} K^{1/(2-\Gamma)} \left[ (\phi^\prime)^2 + \left({4-3\Gamma \over 2-\Gamma}\right)^2 \phi^2 \right] {1 \over R(\theta) \sin^2\theta}. \end{equation} Both the sound speed and the Alfv\`en speed scale as $r^0$ for $\Gamma = 1$, and $r^{1/2}$ for $\Gamma = 0$; for other values of $\Gamma$, the exponent of $r$ lies between these two values. It is also of interest to define the ratio $\mu^2$ of the square of the sound speed and the square of the Alfv\`en speed, each weighted by the density, which is a physical quantity that can be compared with observations: \begin{equation} \mu^2 = {\int_0^{\pi/2} c_s^2\rho\sin\theta d\theta \over \int_0^{\pi/2} v_A^2 \rho\sin\theta d\theta} = 2^{-\Gamma/(2-\Gamma)}{\int_0^{\pi/2} R(\theta)^{\Gamma} \sin\theta d\theta \over \int_0^{\pi/2} \left[ (\phi^\prime)^2 + \left({4-3\Gamma \over 2-\Gamma}\right)^2 \phi^2 \right] /\sin\theta d \theta}. \end{equation} If $c_s$ represents only the thermal sound speed, then the observational summary given by Fuller \& Myers~(1992) would imply that $\mu^2 \sim 1$ in the quiet low-mass cores of GMCs, whereas $\mu^2 \sim 10^{-2}$ in their envelopes. If we include in $c_s$, however, the turbulent contribution, then the turbulent speed is likely to be sub-Alfv\'enic or marginally Alfv\'enic, and $\mu^2 \lesssim 1$ everywhere is probably a better characterization of realistic clouds. \section{RESULTS} \label{sec:results} To obtain an equilibrium configuration for given values of $\Gamma$ and $H_0$, equations (\ref{alongn}) and (\ref{acrossn}) are integrated numerically. The integration is started at $\theta=0$ using the expansions: $\phi = a_0 \xi^2+ \ldots$, $ R = b_0 + b_2 \xi^{4(1-\Gamma)/ (4-3\Gamma)}+ \ldots$, with $\xi=\sin\theta$, and $b_2 = [(4-3\Gamma)/2(1-\Gamma)] A_\Gamma H_0 a_0^{2(1-\Gamma)/(4-3\Gamma)} b_0^{2-\Gamma}$. The values of $a_0$ and $b_0$ are varied until the two BCs at $\theta = \pi/2$ (eq.~\ref{bc}), are satisfied. For flattened equilibria (see below) it is more convenient to start from $\theta=\pi/2$, where the BCs $\phi^\prime (\pi/2)=0$ and $R^\prime(\pi/2)=0$ are imposed, and integrate toward $\theta=0$. The values of $\phi(\pi/2)$ and $R(\pi/2)$ are then varied until a solution is found that satisfies the two BCs at $\theta = 0$. Figure~1 shows the resulting flux and density functions $\phi(\theta)$ and $R(\theta)$ computed for $H_0 = 0.5$ and values of $\Gamma$ between 0.2 and 1. We reproduce the results of \cite{LS96} for $\Gamma = 1$, which is the only case that obtains perfect toroids (i.e., $R[\theta=0] =0$); models with $\Gamma < 1$ have nonzero density at the polar axis. Figure~2 shows the corresponding density contours and magnetic field lines. In the limit $\Gamma\rightarrow 0$, independent of $H_0$ as long as it is nonzero, the pivotal configuration becomes thin disks with an ever increasing magnetic field strength. Table 1 shows the spherical mass to flux ratio $\lambda_r$, the overdensity parameter $D$, and the ratio of the square of the sound and Alfv\`en speeds $\mu^2$. This table shows that, for fixed $H_0$, $\mu^2$ decreases as $\Gamma$ decreases because the magnetic field becomes stronger. For the same reason $D$ increases. In contrast, $\lambda_r$ goes through a minimum as $\Gamma$ decreases. Figures~1 and 2 demonstrate that for $\Gamma \rightarrow 0$ (the logatropic limit), $H_0$ is not a measure of the strength of the magnetic fields since $\phi$ diverges as $R(\theta) \rightarrow \delta(\pi/2-\theta)$ (see \S 5 below). For fixed $\Gamma$, a sequence from small $H_0$ to large $H_0$ progresses through configurations of increasing support by magnetic fields, as demonstrated explicitly for the isothermal case by \cite{LS96}. This behavior is illustrated here for the $\Gamma = 1/2$ case in Figure~3, which shows the density contours and magnetic field lines corresponding to values of $H_0$ from 0.05 to 1.5. Table 2 shows the corresponding values of $\lambda_r$, $D$, and $\mu^2$. For small $H_0$, the equilibria have nearly spherically symmetric isodensity contours and weak quasiuniform magnetic fields that provide little support against gravity. With increasing $H_0$, the pivotal configurations flatten. The case $H_0=1.5$ is already quite disklike: the pole to equator density contrast is $R(\pi/2)/R(0) \simeq 10^6$. For a thin disk, the analysis of \cite{SL97} demonstrates that magnetic tension provides virtually the sole means of horizontal support against self-gravity, with gas and magnetic pressures being important only for the vertical structure. In the limit of a completely flattened disk ($H_0\rightarrow \infty$), $\lambda_r \rightarrow 1$ independent of the detailed nature of the gas EOS (see next section). Table 2 shows the spherical mass to flux ratio $\lambda_r$, the overdensity parameter $D$, and the ratio of the square of the sound and Alfv\`en speeds $\mu^2$. Again $D$ increases monotonically and $\mu^2$ decreases monotonically as the magnetic support increases with $H_0$, while $\lambda_r$ goes through a minimum and tends to 1 for large $H_0$. Since the mass-to-flux ratio $\lambda_r$ is a fundamental quantity that will not change unless magnetic field is lost by ambipolar diffusion, in Figure~4 we consider sequences where $\lambda_r$ is held fixed, but $\Gamma$ is varied. This Figure shows the locus of the set of equilibria with $\lambda_r =0.95,1,$ and $2$ in the ($H_0$, $\Gamma$) plane. Equilibria with $\lambda_r < 1$ are highly flattened when $\Gamma \rightarrow 0$ even for small but fixed values of $H_0$ (see \S 6). In fact, to obtain incompletely flattened clouds when one takes the limit $\Gamma \rightarrow 0$, one also needs simultaneously to consider the limit $H_0\rightarrow 0$. Unfortunately, because both the density and the strength of the magnetic field at the midplane diverge as the equilibria become highly flattened, we are unable to follow numerically the limit $\Gamma \rightarrow 0$ to verify if these sequences of constant $\lambda_r < 1$ will hook to a finite value in the $H_0$ axis, or will loop to $H_0 =0$, consistent with our demonstration in \S4 that flattened disks do not exist in the logatropic limit.\footnote{As the equilibria flatten due to either small $\Gamma$ or large $H_0$, it becomes necessary to determine the constants of the expansions of $R(\theta)$ and $\phi(\theta)$ near the origin with prohibitively increasing accuracy.} We speculate that the results for $\lambda_r < 1$ have the following physical interpretation. According to the theorem of Shu \& Li (1997), only if $\lambda$ itself rather than $\lambda_r$ is less than unity, the magnetic field is strong enough overall to prevent the gravitational collapse of a highly flattened cloud. However, for moderate $H_0$ and $\Gamma$ when $\lambda_r < 1$, even the singular equilibria are probably magnetically subcritical, since there can be little practical difference between the spherical mass-to-flux ratio $\lambda_r$ and the ``true'' mass-to-flux ratio $\lambda$ for highly flattened configurations. The latter is formally infinite when $\Gamma < 1$ only because the mass column goes to zero a little slower than the field column when we perform an integration along the central field line (see footnote 3). In this interpretation, subcritical scale-free clouds with $\lambda_r < 1$ and intermediate values of $\Gamma$ can become highly flattened because magnetic tension supports them laterally against their self-gravity while the soft EOS does not provide much resistance in the direction along the field lines. The squeezing of the cloud toward the midplane is compounded by the {\it confining} pressure of bent magnetic field lines that exert pinch forces in the vertical direction. Both the magnetic tension and the vertical pinch of magnetic pressure disappear when the field lines unbend, as they must to maintain the scale-free equilibria in the limit $\Gamma \rightarrow 0$ (see below). As a consequence, logatropic configurations become spherical for any value of $\lambda_r$. We leave as an interesting problem for future elucidation the determination whether there is still a threshold in $\lambda_r$ below which the SLS, embedded with straight and uniform field lines, will not collapse dynamically. \section{THE THIN DISK LIMIT ($H_0 \gg 1$)} In the limit $H_0\gg 1$, the cloud flattens to a thin disk for any $\Gamma\le 1$. Dominant balance arguments applied to the two ODEs of the problem reveal the following asymptotic behaviour:\footnote{These expansions are not valid for $\Gamma = 1$. See Li \& Shu~(1996) for the correct asymptotic expansion in this case.} \begin{equation} \label{RH0} R(\theta) \rightarrow R_0 \delta(\theta -\pi/2) H_0^{(4-3\Gamma)/(2-\Gamma)} + s(\theta) H_0^{-(4-3\Gamma)/(2-\Gamma)(1-\Gamma)}, \end{equation} \begin{equation} \label{fH0} \phi \rightarrow f(\theta) H_0^{(4-3\Gamma)/(2-\Gamma)}. \end{equation} To the lowest order in $H_0$ the equation of force balance along field lines (eq.~\ref{alongn}) becomes: \begin{eqnarray} \label{asymp} & & {1\over\sin\theta}{d\over d\theta}\left[\sin\theta\left(A_\Gamma f^{-(2-\Gamma)/(4-3\Gamma)} f^{\prime}-s^{-(2-\Gamma)}s^{\prime}\right)\right]= \nonumber \\ & & -2\left[{4-3\Gamma\over (2-\Gamma)^2}s^{-(1-\Gamma)} +\left({4-3\Gamma\over 2-\Gamma}\right)B_\Gamma f^{2(1-\Gamma)/(4-3\Gamma)}\right], \end{eqnarray} valid over the interval $0\leq \theta < \pi/2$, plus the the integral constraint \begin{equation} \label{int1} R_0 -{4-3\Gamma \over (2 - \Gamma)^2} \int_0^\pi s^{-(1-\Gamma)}\sin\theta d\theta - \left({4 - 3 \Gamma \over 2 -\Gamma}\right)^2 B_\Gamma \int_0^\pi f^{2(1-\Gamma)/(4-3\Gamma)} \sin\theta d\theta=0, \end{equation} obtained by integrating eq.~(\ref{alongn}) from $\theta=0 $ to $\pi$, and applying the first BC (eq.~\ref{bc}) on the polar axis. The constant $R_0$ is proportional to the surface density of the polytropic disks, given by \begin{equation} \label{surfden} \Sigma(r) \equiv \lim_{\epsilon\rightarrow 0} \int_{\pi/2-\epsilon}^{\pi/2+\epsilon} \rho r \sin\theta d\theta \rightarrow \left({K \over 2 \pi G}\right)^{1 /(2 - \Gamma)} r^{-\Gamma/(2-\Gamma)} R_0 H_0^{(4-3\Gamma)/(2-\Gamma)}, \end{equation} which, for $\Gamma = 1$ gives $\Sigma \rightarrow H_0 a^2/\pi G r$, as found by \cite{LS96}. Eq.~(\ref{acrossn}) expressing force balance across field lines reduces to \begin{equation} {1\over\sin^2\theta}\left[f^{\prime\prime}-\cot\theta f^\prime+ \ell(\ell +1) f\right]=-C_\Gamma f^{-(2-\Gamma)/(4-3\Gamma)}R_0\delta(\theta-\pi/2), \end{equation} where the parameter $\ell$ is defined by \begin{equation} \ell(\ell +1)\equiv {2(4-3\Gamma) (1-\Gamma)\over (2-\Gamma)^2}. \end{equation} This is equivalent to the equation for force free magnetic fields \begin{equation} \label{forcefree} f^{\prime\prime} - \cot\theta f^\prime + \ell(\ell +1) f = 0, \end{equation} valid over the interval $0\leq \theta < \pi/2$, plus the condition \begin{equation} \label{int2} 2 f^\prime(\pi/2) = C_\Gamma R_0 f(\pi/2)^{-(2-\Gamma)/(4-3\Gamma)}, \end{equation} obtained integrating eq.~(\ref{acrossn}) from $\pi/2 - \epsilon$ to $\pi/2 + \epsilon$, and taking the limit $\epsilon\rightarrow 0$. For integer $\ell$, solutions of eq.~(\ref{forcefree}) regular at $\theta=0$ are Gegenbauer polynomials of order $\ell$ and index $\frac{1}{2}$, $C^{(\frac{1}{2})}_\ell$ (see e.g. Abramowitz \& Stegun~1965). In general, it can be shown (Chandrasekhar~1955) that any axisymmetric force free field, separable in spherical coordinates, can be expressed in terms of fundamental solutions whose radial dependence is given by a combination of Bessel functions of fractional order, and the angular dependence by Gegenbauer polynomials of index $\frac{1}{2}$. In our case, the choice of $\Gamma$ determines a particular exponent of the power-law for the radial part of the flux function, and hence the corresponding value of $\ell$ (non-integer, except for $\Gamma=0$ and 1). Therefore, the magnetic field is force free everywhere except at the midplane where $\rho \neq 0$ and the condition of force balance across field lines has to be satisfied. In the thin disk limit discussed here, the boundary condition $\phi^\prime(\pi/2) =0$ is clearly not fullfilled: the kink of $\phi$ at the midplane provides the magnetic support against self-gravity on the midplane. Currents must exist in the disk to support these kinks. With the definitions $$ y(\theta) \equiv -A_\Gamma {4-3\Gamma\over 2} f(\theta)^{2(1-\Gamma)/(4-3\Gamma)}, ~~~~~ z(\theta)\equiv s(\theta)^{-(1-\Gamma)}, $$ eq.~(\ref{asymp}) transforms into \begin{equation} z^{\prime\prime}+\cot\theta z^{\prime}+\ell(\ell+1)z= y^{\prime\prime}+\cot\theta y^{\prime}+\ell(\ell+1)y, \end{equation} which has the solution $$ z(\theta)=y(\theta)+q(\theta), $$ where $q(\theta)$ is a solution of the homogeneous equation \begin{equation} \label{eqq} q^{\prime\prime}+\cot\theta q^{\prime}+\ell(\ell+1)q=0. \end{equation} Therefore, \begin{equation} \label{sol} s(\theta)=\left[q-A_\Gamma {4-3\Gamma\over 2}f^{2(1-\Gamma)/(4-3\Gamma)} \right]^{-1/(1-\Gamma)}, \end{equation} and the integral constraint eq.~(\ref{int1}) becomes \begin{equation} \label{qconst} \int_0^{\pi/2} q(\theta)\sin\theta d\theta={(2-\Gamma)^2\over 2(4-3\Gamma)}R_0. \end{equation} The problem is thus reduced to the solution of the two homogeneous equations eq.~(\ref{forcefree}) and eq.~(\ref{eqq}) for the functions $f(\theta)$ and $q(\theta)$ which are determined up to an arbitrary constant. However, the two integral constraints that would have determined these latter constants (eqs.~\ref{int2}, \ref{qconst}), contain the additional unknown parameter $R_0$. The system of equations is closed by the requirement that $$ \lim_{H_0 \rightarrow \infty} \lambda_r = 1.$$ Substituting eq.~(\ref{RH0}) and eq.~(\ref{fH0}) in eq.~(\ref{lambdar}), one obtains $$ \lim_{H_0 \rightarrow \infty} \lambda_r = 2^{(1-\Gamma)/(2-\Gamma)} \left({2-\Gamma \over 4-3\Gamma} \right) {R_0 \over 2f({\pi/ 2})} = 1,$$ i.e., \begin{equation} \label{int3} R_0 = 2^{1/(2-\Gamma)} \left( {4-3\Gamma \over 2-\Gamma} \right) f({\pi/ 2}), \end{equation} which gives the remaining condition. Eq.~(\ref{forcefree}) and (\ref{eqq}) can be solved numerically by starting the integration at $\theta=0$ with the series expansions: $q(\theta)=q_0 [ 1-{1 \over 4}\ell(\ell+1)\theta^2 + \ldots ]$, and $f(\theta) = f_0\left\{ \theta^2 -{1 \over 8} [\ell(\ell+1)+{2\over 3}] \theta^4 + \ldots\right\}$, where $q_0$ and $f_0$ are arbitrary constants.\footnote{The two original BCs on the function $R(\theta)$ are of little use here: the one at $\theta=0$ reduces to the condition $\lim_{\theta\rightarrow 0} (1-\Gamma)^{-1}\sin\theta q^\prime=0$, trivially satisfied; the second BC, $R^\prime(\pi/2)=0$ cannot be applied because of the $\delta$-function at $\pi/2$.} The constants $q_0, f_0$ and $R_0$ are then determined by the constraints expressed by eqs.~(\ref{int1}),~(\ref{int2}),and~(\ref{int3}). Figure~5 shows the functions $f(\theta)$ and $s(\theta)$ obtained for $\Gamma=1/2$ and increasing values of $H_0$ from 0.4 to 1.5 compared with the asymptotic expressions computed here. Already for $H_0 = 1.5$, the actual $f(\theta)$ and $s(\theta)$ are very close to the corresponding asymptotic functions eq.~(\ref{RH0}) and eq.~(\ref{fH0}). Table 3 shows the value of the angle $\alpha$ of the magnetic field with the plane of the disk, the flux function $f$ evaluated at $\theta=\pi/2$ (indicative of the magnetic field stength), and the surface density parameter $R_0$, as functions of $\Gamma$. The angle $\alpha$ ranges from $45^\circ$ for the isothermal case $\Gamma = 1$ to $90^\circ$ in the logatropic case $\Gamma =0$. Correspondingly, the magnetic flux in the disk and the surface density both diverge as $\Gamma \rightarrow 0$ for any large but finite value of $H_0$. \section{\bf THE QUASI-SPHERICAL LIMIT ($H_0 \ll 1$)} For the isothermal case $\Gamma=1$, \cite{LS96} have shown how the SIS is recovered for $H_0\ll 1$ from a family of toroids with zero density on the polar axis. For $\Gamma\ne 1$, in the limit $H_0 \ll 1$, the asymptotic expansions are given by: $$R(\theta) \rightarrow \left[ {4-3\Gamma \over (2 -\Gamma)^2} \right]^{1/(2-\Gamma)} + p(\theta) H_0^{(4-3\Gamma)/(3-2\Gamma)} + \ldots$$ $$\phi = g(\theta) H_0^{(4-3\Gamma)/2(3-2\Gamma)} + \ldots.$$ To the lowest order in $H_0$, eqs.~(\ref{alongn}) and (\ref{acrossn}) become: $$ {1\over\sin\theta}{d\over d\theta}\left\{\sin\theta \left[A_\Gamma g^{-(2-\Gamma)/(4- 3\Gamma)} g^\prime- {(2 -\Gamma)^2 \over 4 - 3 \Gamma} p^\prime\right]\right\}= $$ \begin{equation} 2\left[(2-\Gamma)p-\left({4-3\Gamma \over 2-\Gamma}\right)^2 B_\Gamma g^{2(1-\Gamma)/(4-3\Gamma)}\right], \end{equation} and \begin{equation} \label{fluxsmallh} {1\over\sin^2\theta}\left[g^{\prime\prime}-\cot\theta g^\prime+ \ell(\ell+1) g\right]=-C_\Gamma \left[{4-3\Gamma\over (2-\Gamma)^2}\right]^{1/(2-\Gamma)} g^{-(2-\Gamma)/(4-3\Gamma)}, \end{equation} The BC for the functions $p(\theta)$ and $g(\theta)$ are the same as those for $R(\theta)$ and $\phi(\theta)$ in eq.~(\ref{bc}). Figure~6 shows the convergence of the solutions of the full set of equations (\ref{alongn}) and (\ref{acrossn}) obtained for $\Gamma=1/2$ and decreasing values of $H_0$ from 0.4 to 0.05, to the asymptotic solutions obtained by integrating the equations above. Notice that $p(0) < 0$ and $p(\pi/2) > 0$ showing that the sequence of equilibria with $\Gamma=1/2$ originates from the corresponding unmagnetized spherical state (eq.~\ref{nonmagR}) by reducing the density on the pole and enhancing it on the equator. The same behaviour is found for any value of $\Gamma$ in the range $0<\Gamma < 1$. For $\Gamma=1$, the function $p(\theta)$ diverges at $\theta=0$, indicating that this expansion is not appropriate in the isothermal case, as in the case $H_0\gg 1$. For the same reason, the expansion also fails for $\Gamma=0$, since both $g(\theta)$ and $p(\theta)$ diverge on the equatorial plane. These flattened configurations are supported by magnetic and gas pressure against self-gravity. The intensity of the magnetic field can become very high even though $H_0$ is small, because the latter parameter measures not the field strength but the deviations from a force free field (see eq.~\ref{nondim}). \section{\bf THE LOGATROPIC LIMIT ($\Gamma \rightarrow 0$).} We consider in this section the logatropic limit $\Gamma\rightarrow 0$. As anticipated in \S~2, for $\Gamma=0$ eq.~(\ref{alongn}) and (\ref{acrossn}) admit the analytical solution $R=1$ and $\Phi\propto r^2\sin^2\theta$ corresponding to a SLS threaded by a straight and uniform magnetic field. This solution represents the only possible scale-free isopedic configuration of equilibrium for a magnetized cloud with a logatropic EOS. To show this, we use the results of \S~4 and \S~5 for $H_0\gg 1$ and $H_0\ll 1$ to find the limit of the equilibrium configurations for $\Gamma\rightarrow 0$ and fixed (small or large) values of $H_0$. In the limit $H_0 \gg 1$, $\Gamma \rightarrow 0$, analytic solutions to equations ~(\ref{forcefree}) and ~(\ref{eqq}) exist. The magnetic field tends to become uniform and straight, $f(\theta) \rightarrow f(\pi/2)\sin^2\theta$, but $f(\pi/2)$ diverges, as shown in Table~3, and therefore $s(\theta=\pi/2)$ also diverges (see eq.~\ref{sol}). Eq.~(\ref{surfden}) shows in this limit that the surface density $\Sigma$ is independent of $r$, therefore, no pressure gradients can be exerted in the horizontal direction. The value $\Sigma = (K/2 \pi G)^{1/2} R_0 H_0^2 $ diverges as $\Gamma \rightarrow 0$ for any value of $H_0$ because $\lim_{\Gamma \rightarrow 0} R_0 = \infty$ (see Table~3). The magnetic flux threading the disk, $\phi = 2^{-3/2} R_0 H_0^2 r^2 \sin^2\theta$, becomes infinite in order to keep the mass to flux ratio $\lambda_r $ equal to 1. Therefore, the limiting configuration approaches a uniform disk with infinite surface density, threaded by an infinitely strong uniform and straight magnetic field. If we now examine the case $H_0 \ll 1$, in the limit $\Gamma \rightarrow 0$, it is easy to show from eq.~(\ref{fluxsmallh}) that the magnetic field tends to become uniform, $g(\theta) \rightarrow g(\pi/2)\sin^2\theta$, but $\lim_{\Gamma\rightarrow 0}g(\pi/2)=\infty$. Consequently, the density function $p(\theta)$ also diverges in $\theta= \pi/2$, and the configuration again approaches a thin disk threaded by an uniform, infinitely strong, magnetic field. We conclude that scale-free logatropic clouds cannot exist as magnetostatic disks except in some limiting configuration. In the absence of such limits, the equilibria are spherical and can be magnetized only by straight and uniform field lines; i.e., the magnetic field is force-free and therefore given by $H_0=0$. The inside-out gravitational collapse of such a SLS would still proceed self-similarly as in the solution of McLaughlin \& Pudritz (1997), but the frozen-in magnetic fields would yield a dependence with polar angle that eventually produces a pseudodisk (Galli \& Shu 1993a,b; Allen \& Shu 1998a). \section{SUMMARY AND DISCUSSION} We have solved the scale-free equations of magnetostatic equilibrium of isopedic self-gravitating polytropic clouds to find pivotal states that represent the initial state for the onset of dynamical collapse, as first proposed by \cite{LS96} for isothermal clouds. Compared to unmagnetized equilibria, the magnetized configurations are flattened because of magnetic support across field lines. The degree of this support is best represented by the ratio of the square of the sound to Alfv\`en speeds $\mu^2$, or the overdensity parameter $D$, since they are always monotonic functions of $H_0$ and $\Gamma$. Configurations with $\Gamma = 1$ become highly flattened as the parameter $H_0$ increases. When $\Gamma < 1$ (softer EOS) the equilibria get flattened even faster at the same values of $H_0$, since along field lines there is less support from a soft EOS than for a stiff one. However, it seems that in the logatropic limit flattened disks do not exist: the singular scale-free equilibria can only be spherical uniformly magnetized clouds. Figure~7 shows a schematic picture of the $(H_0, \Gamma)$ plane indicating the topology of the solutions for scale free magnetized isopedic singular self-gravitating clouds. In self-gravitating clouds, the joint compression of matter and field is often expressed as producing an expected relationship: $B\propto \rho^\kappa$, with different theorists expressing different preferences for the value of $\kappa$ (e.g., Mestel 1965, Fiedler \& Mouschovias 1993). No local (i.e., point by point) relationship of the form $B\propto \rho^\kappa$ holds for the scale-free equilibria studied in this paper. However, if we average the magnetic field strength and mass density over ever larger spherical volumes centered on $r=0$, we do recover such a relationship: $\langle B \rangle \propto \langle \rho\rangle^\kappa$, where angular brackets denote the result of such a spatial average and $\kappa = \Gamma/2$. We may think of the result $\langle B\rangle \propto \langle \rho\rangle^{\Gamma/2}$ as arising physically from a combination of two tendencies. (a) Slow cloud contraction in the absence of magnetic fields and rotation tends to keep roughly one Jeans mass inside every radius $r$, which yields $\langle \rho\rangle \propto \langle c_s^2\rangle/G r^2$, or $\langle \rho \rangle \propto r^{-2/(2-\Gamma)}$ if $\langle c_s^2\rangle \propto \langle \rho \rangle^{\Gamma -1}$. (b) Slow cloud contraction in the absence of gas pressure tends to keep roughly one magnetic critical mass inside every radius $r$, which yields $\langle B\rangle /r\langle \rho \rangle \propto \lambda_r$ = constant, or $\langle B \rangle \propto r\langle\rho\rangle \propto r^{-\Gamma/(2-\Gamma)} \propto \langle \rho\rangle^{\Gamma/2}$ if gas pressure (thermal or turbulent) plays a comparable role to magnetic fields in cloud support. Notice that our reasoning does not rely on arguments of cloud geometry, e.g., whether cloud cores flatten dramatically or not as they contract; nor does it depend sensitively on the precise reason for core contraction, e.g., because of ambipolar diffusion or turbulent decay. Crutcher (1998) claims that the observational data are consistent with $\kappa = 0.47 \pm 0.05$. If we take Crutcher's conclusion at face value, we would interpret the observations as referring mostly to regions where the EOS is close to being isothermal $\Gamma \approx 1$, which is the approximation adopted by many theoretical studies that ignore the role of cloud turbulence. The result is not unexpected for low-mass cloud cores, but we would not naively have expected this relationship for high-mass cores and cloud envelopes, where the importance of turbulent motions is much greater. Unfortunately, the observational data refer to different clouds rather than to different (spatially averaged) regions of the same cloud, so there is some ambiguity how to make the proper connection to different theoretical predictions. There may also be other mechanisms at work, e.g., perhaps a tendency for observations to select for regions of nearly constant Alfv\'en speed, $\langle v_A \rangle \propto \langle B\rangle/\langle \rho \rangle^{1/2} \approx$ constant (Bertoldi \& McKee 1992). Thus, we would warn the reader against drawing premature conclusions about the effective EOS for molecular clouds, or the related degree to which observations can at present distinguish whether molecular clouds are magnetically supercritical or subcritical. If molecular clouds are magnetically supercritical, with $\lambda_r$ greater than 1 by order unity (say, $\lambda_r = 2$), then an appreciable fraction (say, 1/2) of their support against self-gravity has to come from turbulent or thermal pressure (Elmegreen 1978, McKee et al. 1993, Crutcher 1998). Modeled as scale-free equilibria, such clouds with $\mu^2$ of order unity are not highly flattened (see Tables 1, 2 and Figs. 2, 3). Suppose we try gravitationally to extract a subunit from an unflattened massive molecular cloud, where the cloud as a whole is only somewhat supercritical, $\lambda_r \sim 2$. If the subunit's linear size is smaller than the vertical dimension of the cloud by more than a factor of 2, which will be the case if we consider subunits of stellar mass scales, then this subunit will not itself be magnetically supercritical. Magnetically subcritical pieces of clouds cannot contract indefinitely without flux loss, so star formation in {\it unflattened} clouds, if they are not highly supercritical, needs to invoke some degree of ambipolar diffusion in order to produce small dense cores that can gravitationally separate from their surroundings. On the other hand, if molecular clouds are magnetically critical or subcritical, with $\lambda_r \le 1$, then almost all scale-free equilibria are highly flattened, with $\mu^2$ appreciably less than unity. On a small scale, any subunit of this cloud, even subunits with vertical dimension comparable to the cloud as a whole, would also be magnetically critical or subcritical. For such a subunit to contract indefinitely, we would again need to invoke ambipolar diffusion to make a cloud core magnetically supercritical. Thus, although the decay of turbulence can accelerate the formation of cloud cores, the ultimate formation of stars from such cores may still need to rely on {\it some} magnetic flux loss (but perhaps not more than a factor of $\sim 2$) to trigger the evolution of the cores toward gravomagneto catastrophe and a pivotal state with a formally infinite central concentration. On the large scale, if GMCs are modeled as flattened isopedic sheets, \cite{SL97} proved that magnetic pressure and tension are proportional to the gas pressure and force of self-gravity. Their theorems hold independently of the detailed forms of the EOS or the surface density distribution. If GMCs are truly highly flattened -- with typical dimensions, say, of 50 pc $\times$ 50 pc $\times$ a few pc or even less -- then many aspects of their magnetohydrodynamic stability and evolution become amenable to a simplified analysis through the judicious application and extension of the theorems proved by \cite{SL97} (e.g., see Allen \& Shu 1998b, \cite{shuetal99}). This exciting possibility deserves further exploration. \bigskip \acknowledgments D.G. acknowledges support by CNR grant 97.00018.CT02, ASI grant ARS-96-66 and ARS-98-116, and hospitality from UNAM, M\'exico. S.L. acknowledges support by J. S. Guggenheim Memorial Foundation, grant DGAPA-UNAM and CONACyT, and hospitality from Osservatorio di Arcetri. F.C.A. is supported by NASA grant No.~NAG~5-2869 and by funds from the Physics Department at the University of Michigan. The work of F.H.S. is supported in part by an NSF grant and in part by a NASA theory grant awarded to the Center for Star Formation Studies, a consortium of the University of California at Berkeley, the University of California at Santa Cruz, and the NASA Ames Research Center. \clearpage
\section{INTRODUCTION} \label{sect1} Stable mass transfer on to a (compact) accretor which is driven by the nuclear evolution of a giant is a process of considerable astrophysical interest. This type of evolution was first encountered in numerical simulations by (among others) Kippenhahn, Kohl \& Weigert (1967) and then studied in some detail by Refsdal \& Weigert (1969, 1971) in the context of the formation of low-mass white dwarfs in binaries. Later Webbink, Rappaport \& Savonije (1983, hereafter WRS) and, independently, Taam (1983) have proposed that some of the most luminous among the galactic low-mass X-ray binaries could be powered by nuclear time-scale driven mass transfer from a giant. Finally with the detection of the millisecond pulsar binary PSR 1953+29 by Boriakoff, Buccheri \& Fauci (1983) it became immediately clear that this system (and many other similar ones which have been found since then) must be the end product of an evolution of the type described by WRS and Taam (1983), see e.g. Joss \& Rappaport (1983), Paczynski (1983), Savonije (1983), and Rappaport et al. (1995, hereafter RPJDH) for a more recent reference. Whereas Taam (1983) performed full stellar structure calculations WRS took a simpler approach by making use of the well-known fact that giants obey a core mass-luminosity and a core mass-radius relation. WRS derived these relations from full stellar equilibrium models of giants and approximated them as power series of the mass of the degenerate helium core $M_{\rm c}$. By using the core mass-luminosity and core mass-radius relation calculating the evolution of a binary with nuclear time-scale driven mass transfer from a giant reduces to preforming a simple integral in time which, however, in the approach taken by WRS has still to be performed numerically. The purpose of this paper is first to show that the evolution of such a binary can be described by an analytical solution if the core mass-luminosity and core mass-radius relation are simple power laws of $M_{\rm c}$, and second to show that the results of full stellar structure calculations of giants given by WRS can indeed be approximated with little loss of accuracy by simple power laws. Moreover, the analytical solutions which we present are more general than the numerical ones of WRS in the sense that we allow for non-conservative mass transfer and a non-zero mass radius exponent of the donor star (see below section \ref{sect3}). The advantages of having an analytical solution as compared to a numerical one are immediatedly clear: First, if sufficiently simple (as in our case), such solutions provide us with direct physical insight into how this type of evolution works, i.e. which parameters determine the characteristc properties and which ones are of minor importance. Second, whereas a numerical solution has to be computed separately for each set of input parameters of interest, an analytical solution provides us with the whole manifold of solutions, i.e. we get the explicit dependence of the solution on the input parameters. Among other things, this allows us to examine whether certain properties found in numerical solutions are generic or a simple consequence of a particular choice of parameters. A case in point is the well-known property of the solutions presented by WRS, namely that the mass transfer rate stays almost constant over most of the mass transfer phase. Another area of application in which simple analytical solutions are superior to numerical ones is population synthesis. This is because population synthesis calculations usually involve integrals over the complete manifold of solutions of the particular evolutionary phase under consideration. Using numerical solutions for that task requires computation of a sufficiently dense grid of evolutions over a multi-dimensional parameter space and such computations can easily become prohibitively expensive (see e.g. Kolb 1993 for a detailed discussion in the context of cataclysmic variables). Using analytical solutions instead is comparatively easy and cheap. This paper is organized as follows: In section \ref{sect2} we present the basic model assumptions in the framework of which we shall work, in section \ref{sect3} we give an analytical solution for the nuclear evolution of a giant, and in section \ref{sect4} we derive expressions for the mass radius exponent of the donor star's critical Roche radius. Finally, in section \ref{sect5} we present the analytical solution for nuclear time-scale driven mass transfer from a giant. In order to compare our results with those of WRS we derive in section \ref{sect6} the parameters of the power law approximations of the core mass-luminosity and core mass-radius relations using the data from WRS and RPJDH. In section \ref{sect7} we then compare the results obtained from the analytical solutions with the corresponding numerical ones of WRS. In section \ref{sect8} we examine whether the model is selfconsistent and whether it is applicable to low-mass X-ray binaries. A discussion based on the analytical solution of the key features of nuclear time-scale driven mass transfer from a giant follows in section \ref{sect9}, and our main conclusions are summarized in the final section \ref{sect10}. \section{BASIC MODEL ASSUMPTIONS} \label{sect2} In the following we shall list the basic model assumptions in the framework of which we shall work. These assumptions are: \begin{enumerate} \item The donor star (of mass $M_2$) is on the first giant branch and has a degenerate helium core of mass $M_{\rm c}$. Nuclear luminosity comes from hydrogen shell burning exclusively. \item $M_2$ is small enough that the binary is adiabatically and thermally stable against mass transfer. For this to hold we must have (e.g. Ritter 1988) \begin{equation} \zeta_{\rm{ad}} - \zeta_{\rm{R,2}} > 0 \label{2.1} \end{equation} and \begin{equation} \zeta_{\rm{e}} - \zeta_{\rm{R,2}} > 0, \label{2.2} \end{equation} where \begin{equation} \zeta_{\rm{ad}} = {\biggl(\frac{\partial \ln R_2}{\partial \ln M_2} \biggr)}_{\rm{ad}} \label{2.3} \end{equation} is the adiabatic mass radius exponent of the donor star, \begin{equation} \zeta_{\rm{e}} = {\biggl(\frac{\partial \ln R_2}{\partial \ln M_2} \biggr)}_{\rm{e}} \label{2.4} \end{equation} the corresponding thermal equilibrium mass radius exponent, and \begin{equation} \zeta_{\rm{R,2}} = {\biggl(\frac{\partial \ln R_{R,2}}{\partial \ln M_2} \biggr)}_{\ast} \label{2.5} \end{equation} the mass radius exponent of its critical Roche radius (see e.g. Ritter 1996 for details). In (\ref{2.5}) the subscript $\ast$ indicates that for computing this quantity one has to specify how, as a consequence of mass transfer, mass and orbital angular momentum are lost from and redistributed within the binary system (see below section \ref{sect4}). For stars on the first giant branch and with small relative core mass $M_{\rm c}/M$ $\zeta_{\rm{ad}} \approx -1/3$. $\zeta_{\rm{ad}}$ increases with relative core mass and becomes positive if $M_{\rm{c}}/M \ga 0.2$ (Hjellming \& Webbink 1987). Therefore, as we shall see, adiabatic stability of mass transfer is less critical than thermal stability. Assuming that the donor obeys a core mass-luminosity relation, i.e. that $L = L(M_{\rm c})$ and $\partial L/\partial M = 0$, where $L$ is the (nuclear) luminosity, $\zeta_{\rm e}$ follows from the Stefan--Boltzmann law: \begin{equation} \zeta_{\rm e} = \frac{1}{2}\, {\left(\frac{\partial \ln L}{\partial \ln M}\right)}_{\rm e} - 2\, \left(\frac{\partial \ln T_{\rm eff}}{\partial \ln M}\right)_{\rm e} = -2\, {\left(\frac{\partial \ln T_{\rm eff}}{\partial \ln M}\right)}_{\rm e} \label{2.6} \end{equation} It is well known that a giant, as long as it has a sufficiently deep outer convective envelope, in the HRD stays close to and evolves along the Hayashi line to higher luminosity. This property allows us to get an approximation for $\zeta_{\rm e}$. Approximating the Hayashi line in the HRD in the form (e.g. Kippenhahn \& Weigert 1990) \begin{equation} \log L = a \log T_{\rm eff} + b \log M + c \label{2.7} \end{equation} this together with (\ref{2.6}) yields \begin{equation} \zeta_{\rm e} \approx - 2 b/ a\,. \label{2.8} \end{equation} Numerical calculations typically yield $-0.3 \la \zeta_{\rm e} \la -0.2$. Thus, except for small relative core mass, $\zeta_{\rm e} < \zeta_{\rm ad}$, i.e. the criterion for thermal stability (\ref{2.2}) against mass transfer is stronger than the one for adiabatic stability (\ref{2.1}). \item Mass transfer is driven only by the donor's nuclear evolution, i.e. we assume that there is no systemic angular momentum loss in the absence of mass transfer. Consequential angular momentum loss (see e.g. King \& Kolb 1995 for a discussion) is, however, not excluded a priori. \item Mass transfer is so slow that the donor star remains close to thermal equilibrium. Accordingly we assume that its radius $R_2$ and luminosity $L_2$ are given by the corresponding thermal equlilibrium values, i.e. $R_2 = R_{2,{\rm e}}$ and $L_2 = L_{2,{\rm e}}$, and hence $\zeta_{\rm eff} = (d \ln R_2/d \ln M_2) = \zeta_{\rm e}$. This condition is well fulfilled because mass transfer occurs on the nuclear time scale $\tau_{\rm nuc}$ which is much longer than the Kelvin-Helmholtz time $\tau_{\rm KH} = G {M_2}^2/R_2 L_2$. We shall later (section \ref{sect8}) verify the validity of this assumption. \end{enumerate} With the above assumptions we can now write the mass transfer rate, i.e. the donor's mass loss rate as (e.g. Ritter 1996) \begin{equation} -\dot{M}_2 = \frac{M_2}{\zeta_{\rm e} - \zeta_{\rm R,2}} \left(\frac{\partial \ln R_2}{\partial t}\right)_{\rm nuc}, \label{2.9} \end{equation} where $(\partial \ln R_2/\partial t)_{\rm nuc}$ is the inverse time-scale of the donor's expansion due to nuclear evolution. Next we shall work out in more detail $(\partial \ln R_2/ \partial t)_{\rm nuc}$ and $\zeta_{\rm R,2}$. \section{NUCLEAR EVOLUTION OF A GIANT} \label{sect3} Here we are going to exploit the fact that to a very good approximation the luminosity $L$ of a star on the first giant branch scales as a simple power law of the mass $M_{\rm c}$ of the degenerate helium core but does not depend on its total mass (Refsdal \& Weigert 1970). Accordingly we make the following ansatz for $L$: \begin{equation} L(M_{\rm c}) = L_0 {\left(\frac{M_{\rm c}}{{\rm M_{\odot}}}\right)}^\lambda \label{3.1} \end{equation} It then follows from Eqs. (\ref{2.6}) and (\ref{2.7}) that the radius of such a star cannot at the same time depend only on the core mass $M_{\rm c}$ but must also depend on its total mass $M$. Because along the Hayashi line the effective temperature remains almost constant it follows from (\ref{3.1}) and the Stefan--Boltzmann law that the appropriate form of the `core mass-radius relation' is \begin{equation} R(M_{\rm c}, M) = R_0 {\left(\frac{M_{\rm c}}{{\rm M_{\odot}}}\right)}^\rho {\left(\frac{M}{{\rm M_{\odot}}}\right)}^{\zeta_{\rm e}}. \label{3.2} \end{equation} Here, $L_0$, $\lambda$, $R_0$ and $\rho$ are parameters which have to be determined from full numerical calculations. We shall give examples (taken form WRS and RPJDH) in section \ref{sect6}. Using homology techniques Refsdal \& Weigert (1970) could derive an expression for $\lambda$, the numerical value of which depends on the conditions in the hydrogen shell source. A typical value for Pop. I giants is $\lambda \approx 7 - 8$. Hydrogen burning in the shell source around the degenerate core adds to the core mass at a rate $\dot M_{\rm c}$. Thereby the nuclear luminosity generated is \begin{equation} L =L_{\rm nuc} = X Q \dot M_{\rm c}\,, \label{3.3} \end{equation} where $X$ is the hydrogen mass fraction in the star's envelope and $Q \approx 6~10^{18}$ erg g$^{-1}$ the net energy yield of hydrogen burning per unit mass. In writing (\ref{3.3}) we have assumed that the star is in thermal equilibrium, i.e. that $L =L_{\rm nuc}$ or, for practical purposes, that $|L_g| \ll L_{\rm nuc}$, where $L_g$ is the gravo-thermal luminosity. (\ref{3.1}) and (\ref{3.3}) together yield a simple differential equation for $M_{\rm c}(t)$ with the solution \begin{equation} M_{\rm c}(t) = M_{\rm c}(t=0) \left(1 - \frac{t}{t_\infty} \right)^{1\over{1 - \lambda}}, \label{3.4} \end{equation} where \begin{equation} t_{\infty} = \frac{X Q {\rm M_{\odot}}}{(\lambda - 1) L_0} {\left[\frac {M_{\rm c}(t=0)}{{\rm M_{\odot}}}\right]}^{1-\lambda} \label{3.5} \end{equation} is the time over which, formally, the core mass reaches infinity. It is also the characteristic time for nuclear evolution on the first giant branch. (\ref{3.4}) combined with respectively (\ref{3.1}) or (\ref{3.2}) yields \begin{equation} L(t) = L_0\, {\left[ \frac{M_{\rm c}(t=0)}{{\rm M_{\odot}}}\right]}^\lambda {\left(1 - \frac{t}{t_\infty} \right)}^{\frac{\lambda} {1 - \lambda}} \label{3.6} \end{equation} and \begin{equation} R(t) = R_0\, {\left(\frac{M}{{\rm M_{\odot}}}\right)}^{\zeta_{\rm e}} {\left[ \frac{M_{\rm c}(t=0)}{{\rm M_{\odot}}}\right]}^\rho {\left(1 - \frac{t}{t_\infty} \right)}^{\frac{\rho}{1 - \lambda}}. \label{3.7} \end{equation} Here $M_{\rm c}(t=0)$ is the core mass at the beginning of the evolution on the first giant branch. This quantity has also to be taken from full stellar evolution calculations or to be treated as a free parameter. From (\ref{3.7}) we finally obtain \begin{equation} \left(\frac{\partial \ln R}{\partial t}\right)_{\rm nuc} = \frac{\rho L_0}{X Q {\rm M_{\odot}}} {\left(\frac{M_{\rm c}} {{\rm M_{\odot}}}\right)}^{\lambda -1} = \frac{\rho}{\lambda - 1}~ \frac{1}{t_\infty - t}\,, \label{3.8} \end{equation} the expression to be inserted in (\ref{2.9}). \section{COMPUTING $\zeta_{\rm R,2}$} \label{sect4} Here we follow standard arguments which have been given many times in the literature (e.g. Soberman, Phinney \& van den Heuvel 1997 and references therein). Therefore, we shall give here only the definitions we use and the final results. In the context of this paper, we assume that mass transfer occurs from the secondary and that the primary (of mass $M_1$) accretes (on average) a fraction $\eta$ of the transferred mass, i.e. \begin{equation} dM_1 = - \eta dM_2 \label{4.1} \end{equation} and that the mass $d(M_1 + dM_2) = (1 - \eta) dM_2$ leaves the binary system carrying away the orbital angular momentum \begin{equation} dJ = \nu J \frac{d(M_1 + M_2)}{M_1 + M_2}\,, \label{4.2} \end{equation} where, for the moment at least, $\eta$ and $\nu$ are free parameters, \begin{equation} J = {\left( \frac{G {M_1}^2 {M_2}^2 a}{M_1 + M_2}\right)}^{\frac{1}{2}} \label{4.3.} \end{equation} is the orbital angular momentum, and $a$ the orbital separation. Writing for the secondary's critical Roche radius \begin{equation} R_{\rm R,2} = a f_2(q), \label{4.4} \end{equation} where \begin{equation} q = \frac{M_1}{M_2} \label{4.5} \end{equation} is the mass ratio, one derives \begin{equation} \zeta_{\rm R,2} = (1 - \eta) \frac{2 \nu + 1}{1 + q} + \frac{2 \eta}{q} - 2 - {\beta}_2 \left(1 + \frac{\eta}{q} \right), \label{4.6} \end{equation} with \begin{equation} {\beta}_2 = \frac{d \ln f_2}{d \ln q}\,. \label{4.7} \end{equation} As a further simplification we use for $f_2$ in (\ref{4.4}) and (\ref{4.7}) the approximation (Paczynski 1971) \begin{equation} {f_2}(q) = {\left(\frac{8}{81}\right)}^{1/3} (1 + q)^{-1/3}, ~~~q \ga 1.25 \label{4.8} \end{equation} which yields \begin{equation} {\beta}_{2}(q) = - \frac{q}{3 (1 + q)}\,. \label{4.9} \end{equation} Inserting (\ref{4.9}) in (\ref{4.6}) finally yields \begin{equation} \zeta_{\rm R,2} = \frac{(1 - \eta)(2 \nu + 1)}{1 + q} + \frac{2 \nu}{q} -2 + \frac{\eta + q}{3(1 + q)}\,. \label{4.10} \end{equation} In the following we shall also give the results for two specific choices of the parameters $\eta$ and $\nu$: \subsection{Conservative mass transfer} \label{sect4.1} Here $\eta = 1$ and $\nu = 0$. From (\ref{4.6}) and (\ref{4.9}) we then have \begin{equation} \zeta_{\rm R,2} = \frac{6 - 5 q}{3 q}\,. \label{4.12} \end{equation} \subsection{Isotropic stellar wind from the primary} \label{sect4.2} Here we are considering the case that all the transferred mass is ejected from the primary in an isotropic stellar wind carrying with it the specific orbital angular momentum of the primary. Accordingly $\eta =0$ and $\nu =1/q$. Inserted in (\ref{4.6}) together with (\ref{4.9}) this yields \begin{equation} \zeta_{\rm R,2} = \frac{6 - 5q^2 - 3q}{3q(1 + q)}\,. \label{4.14} \end{equation} \section{EVOLUTION WITH MASS TRANSFER FROM A GIANT} \label{sect5} Now that we have determined $(\partial \ln R/\partial t)_{\rm nuc}$ and $\zeta_{\rm R,2}$ in equation (\ref{2.9}) we can solve that equation. First we are going to give a rather general solution before discussing special cases of interest. \subsection{General analytical solution} \label{sect5.1} General solution in this context means that we allow for a constant value $\eta \ge 0$, either a constant value of $\nu \ge 0$ or $\nu = 1/q$, and a constant value of $\zeta_{\rm e}$ subject to the condition (\ref{2.2}). Inserting (\ref{3.8}) and (\ref{4.6}) with (\ref{4.9}) in (\ref{2.9}) yields an ordinary differential equation for ${M_2}(t)$, or rather, as we shall see, for $t(M_2)$ which is solved by separating the variables $M_2$ and $t$: \begin{equation} \int^{M_2}_{M_{\rm 2,i}}\frac{\zeta_{\rm e}-\zeta_{\rm R,2}}{m_2}\,dm_2 = \int^t_0 {\left(\frac{\partial \ln R}{\partial \tau}\right)}_{\rm nuc}\,d \tau \label{5.1} \end{equation} The solution of (\ref{5.1}) can be written as follows: \begin{equation} t(M_2)=t_{\infty}\,{\left[1-{\left(\frac{M_1}{M_{\rm 1,i}}\right)}^{-p_1} {\left(\frac{M_2}{M_{\rm 2,i}}\right)}^{-p_2} {\left(\frac{M_1 + M_2}{M_{\rm 1,i} + M_{\rm 2,i}}\right)}^{-p_3} e^{-p_4\, \frac{M_{\rm 2,i}-M_2}{M_1}} \right]}\,, \label{5.2} \end{equation} where the exponents $p_1$, $p_2$, $p_3$ and $p_4$ are given in Table 1 for various cases of interest. With (\ref{5.2}) the mass transfer rate can be written as \begin{equation} -\dot M_2 = \frac{1}{\zeta_{\rm e}-\zeta_{\rm R,2}}\, \frac{\rho L_0}{X Q}\, {\left(\frac{M_{\rm c,i}}{{\rm M_{\odot}}}\right)}^{\lambda -1}\, \left(\frac{M_2}{{\rm M_{\odot}}}\right)\, {\left(1-\frac{t}{t_{\infty}} \right)}^{-1}\,. \label{5.3} \end{equation} Inserting (\ref{5.2}) in (\ref{3.6}) and (\ref{3.7}) finally also yields \begin{equation} L_2 = L_0\,{\left(\frac{M_{\rm c,i}}{{\rm M_{\odot}}}\right)}^{\lambda}\, \left[{\left(\frac{M_1}{M_{\rm 1,i}}\right)}^{p_1}\, {\left(\frac{M_2}{M_{\rm 2,i}}\right)}^{p_2}\, {\left(\frac{M_1 + M_2}{M_{\rm 1,i} + M_{\rm 2,i}}\right)}^{p_3}\, e^{p_4\,\frac{M_{\rm 2,i} - M_2}{M_1}}\right]^ {\frac{\lambda}{\lambda -1}} \label{5.5} \end{equation} and \begin{equation} R_2 = R_0\,{\left(\frac{M_{\rm c,i}}{{\rm M_{\odot}}}\right)}^{\rho}\, \left(\frac{M_2}{{\rm M_{\odot}}}\right)^{\zeta_{\rm e}}\, {\left[{\left(\frac{M_1}{M_{\rm 1,i}}\right)}^{p_1}\, {\left(\frac{M_2}{M_{\rm 2,i}}\right)}^{p_2}\, {\left(\frac{M_1 + M_2}{M_{\rm 1,i} + M_{\rm 2,i}}\right)}^{p_3}\, e^{p_4\,\frac{M_{\rm 2,i} - M_2}{M_1}}\right]}^ {\frac{\rho}{\lambda -1}}\,. \label{5.6} \end{equation} Here $M_{\rm 1,i}$, $M_{\rm 2,i}$ and $M_{\rm c,i}$ are respectively the initial values, i.e. taken at $t = 0$, of $M_1$, $M_2$, and $M_{\rm c}$. Note that from (\ref{4.1}) we have \begin{equation} M_1 = M_{\rm 1,i} + \eta \left(M_{\rm 2,i} - M_2 \right)\,. \label{5.7} \end{equation} As can be seen the equations (\ref{5.3}) (together with (\ref{5.2})), (\ref{5.5}) and (\ref{5.6}) do not depend explicitly on time $t$. The solutions are completely fixed by the parameters of the core mass-luminosity relation (\ref{3.1}), i.e. $L_0$ and $\lambda$, the `core mass-radius relation' (\ref{3.2}), i.e. $R_0$, $\rho$ and $\zeta_{\rm e}$, the initial values and the current values of both masses and the secondary's core mass. Kowing $\dot M_2$, $R_2$ and $L_2$ as explicit functions of the initial values and current values of $M_1$, $M_2$ and $M_{\rm c}$ other quantities of interest such as the orbital period $P$, the orbital separation $a$, the size of the accretor's Roche radius $R_{\rm R,1}$ and with it the size of the disc around the accretor and the secondary's effective temperature $T_{\rm eff,2}$ follow directly. What this solution does not provide explicitly is the total duration of the mass transfer phase., i.e. the time $t_{\rm f}$ when the secondary's envelope is exhausted or, which is equivalent, the final core mass $M_{\rm c,f}$. A very good estimate of the latter together with $t_{\rm f}$ can be obtained by inserting (\ref{5.2}) with $M_2 = M_{\rm 2,f} \approx M_{\rm c,f}$ in (\ref{3.4}). Note that the resulting equation for $M_{\rm c,f}$ cannot in general be solved in closed form. Reinserting the solution of this equation in (\ref{5.2}) finally yields also $t_{\rm f}$. \subsection{Special cases} \label{sect5.2} Table 1 lists the powers $p_1$, $p_2$, $p_3$ and $p_4$ which appear in equations (\ref{5.2}) -- (\ref{5.6}) for a number of special cases of interest. Case 1 is the most general case for constant values of $\eta$, $\nu$ and $\zeta_{\rm e}$. In case 2 the special choice $\nu = 1/q$ with $\eta > 0$ is made, whereas case 3 corresponds to case 2 in the limit $\eta = 0$. Cases 4 and 5 correspond to conservative mass transfer with respectively $\zeta_{\rm e} = {\rm const.} \ne 0$ and $\zeta_{\rm e} = 0$. Case 6 finally corresponds to case 3 with $\zeta_{\rm e} = 0$. Case 3 (or 6) corresponds to the situation in which the accretor ejects all the transferred mass with the specific orbital angular momentum of the accretor's orbit. This approximates what apparently must happen in systems in which the mass transfer rate to a neutron star exceeds the Eddington accretion rate by many orders of magnitude (see e.g. King \& Ritter 1999 for a discussion). We shall later mostly make use of case 5 because this is the approximation within which WRS have performed their numerical calculations and with which we wish to compare the results of our analytical solution. Before we can do that we have to provide the parameters describing the core mass-luminosity and core mass-radius relation. \begin{table*} \begin{minipage}{\hsize} \caption{Parameters and coefficients entering the analytical solution} \begin{tabular}{rccccccc}\hline case & $\eta$ & $\nu$ &$\zeta_{\rm e}$& $p_1$ & $p_2$ & $p_3$ & $p_4$ \\ \hline & & & & & & & \\ 1 &$\ge 0$, const.& const.& const. &$-\frac{2(\lambda -1)}{\rho}$ &$-\frac{(3\zeta_{\rm e}+5)(\lambda -1)}{3\rho}$&$\frac{2(3\nu +1)(\lambda -1)}{3\rho}$& 0 \\ & & & & & & & \\ 2 &$ > 0$, const.&$1/q$ & const. &$-\frac{2(\lambda -1)}{\eta \rho}$&$-\frac{(3\zeta_{\rm e}+5)(\lambda -1)}{3\rho}$&$-\frac{4(\lambda -1)}{3\rho}$ & 0 \\ & & & & & & & \\ 3 & 0 &$1/q$ & const. & 0 &$-\frac{(3\zeta_{\rm e}+5)(\lambda -1)}{3\rho}$&$-\frac{4(\lambda -1)}{3\rho}$ &$-\frac{2(\lambda -1)}{\rho}$\\ & & & & & & & \\ 4 & 1 & ---- & const. &$-\frac{2(\lambda -1)}{\rho}$ &$-\frac{(3\zeta_{\rm e}+5)(\lambda -1)}{3\rho}$& 0 & 0 \\ & & & & & & & \\ 5 & 1 & ---- & 0 &$-\frac{2(\lambda -1)}{\rho}$ &$-\frac{5(\lambda -1)}{3\rho}$ & 0 & 0 \\ & & & & & & & \\ 6 & 0 &$1/q$ & 0 & 0 &$-\frac{5(\lambda -1)}{3\rho}$ &$-\frac{4(\lambda -1)}{3\rho}$ &$-\frac{2(\lambda -1)}{\rho}$\\ \hline \end{tabular} \end{minipage} \end{table*} \section{Approximations for the core mass-luminosity and core mass radius relation} \label{sect6} In general the values of $L_0$, $\lambda$, $R_0$, $\rho$ and $\zeta_{\rm e}$ in equations (\ref{3.1}) and (\ref{3.2}) have to be determined from full numerical calculations of stars in thermal equilibrium on the first giant branch. Here we restrict ourselves to deriving approximate values of these parameters by using results given by WRS (their equations 4 -- 6). Because WRS have assumed $\zeta_{\rm e} = 0$ their relation (5) is a true core mass-radius relation in the sense that the radius of the star does not depend on the total mass $M$. Their results for Pop. I and Pop. II giants are shown graphically respectively in Fig. 1 and Fig. 2. It is immediately apparent that in the range $0.15 {\rm M_{\odot}} \la M_{\rm c} \la 0.4 {\rm M_{\odot}}$ the relations $L(M_{\rm c})$ and $R(M_{\rm c})$ are very nearly simple power laws in $M_{\rm c}$. Based on the results of Refsdal \& Weigert (1970) this is of course not surprising. Because our analytical solutions of section \ref{sect5} are based on the power law approximations (\ref{3.1}) and (\ref{3.2}) we fit the results of WRS shown in Figs. 1 and 2 accordingly. The resulting parameters, together with the chemical composition of the underlying numerical models (characterized in the usual way by the hydrogen mass fraction $X$ and the the metallicity $Z$) are summarized in Table 2. The corresponding relations are shown as thin full ($L$) or dashed ($R$) lines in Figs. 1 and 2. \begin{figure \begin{minipage}[t]{0.48\hsize} \centerline{\epsfxsize=1.0\hsize\epsffile{Fig1.eps}} \caption{Core mass-luminosity relation (full line), core mass-radius relation (long-dashed line) and core mass-effective temperature relation (short-dashed line) for Pop. I giants according to numerical calculations by WRS. Our simple power law fits are shown respectively as thin full and thin long-dashed lines. The corresponding fit parameters are listed in Table 2.} \end{minipage} \hfill \begin{minipage}[t]{0.48\hsize} \centerline{\epsfxsize=1.0\hsize\epsffile{Fig2.eps}} \caption{The same as Fig. 1, but for Pop. II giants.} \end{minipage} \end{figure} For comparison we provide in Table 2 also approximate parameters of a power law fit in the range $0.2 {\rm M_{\odot}} \la M_{\rm c} \la 0.5 {\rm M_{\odot}}$ to the core mass-radius relation derived by RPJDH (their equation 5). \begin{table*} \begin{minipage}{\hsize} \caption{Approximate parameters of the core mass-luminosity and core mass-radius relations derived from numerical results} \begin{tabular}{ccccccccc}\hline Pop. & X & Z & $\log\frac{L_0}{{\rm L_{\odot}}}$ & $\lambda$ & $\log\frac{R_0}{{\rm R_{\odot}}}$ & $\rho $ & $\zeta_{\rm e}$ & source \\ \hline & & & & & & & & \\ I & 0.70 & 0.02 & 6.3 & 8 & 4.1 & 5 & 0 & WRS \\ & & & & & & & & \\ II & 0.80 &0.0001 & 4.5 & 5 & 2.7 & 3 & 0 & WRS \\ & & & & & & & & \\ I & 0.70 & 0.02 & --- & --- & 3.58 & 4.21 & 0 & RPJDH \\ & & & & & & & & \\ II & 0.70 & 0.001 & --- & --- & 3.36 & 4.17 & 0 & RPJDH \\ \hline \end{tabular} \end{minipage} \end{table*} \section{NUMERICAL RESULTS VERSUS THE ANALYTICAL MODEL} \label{sect7} Here we wish to compare the results obtained from the analytical approximation, i.e. from equations (\ref{5.2}) -- (\ref{5.6}), with the numerical results obtained by WRS. For this we have to insert the parameters listed in Table 2 into the expressions for $p_1$, $p_2$, $p_3$, and $p_4$ corresponding to case 5 in Table 1 and in equations (\ref{5.2}) -- (\ref{5.6}). Rather than listing all the resulting equations for both Pop. I and Pop. II composition, we restrict ourselves to giving only the corresponding expressions for the mass transfer rate. With $Q = 6~10^{18}$ erg g$^{-1}$ this yields \begin{equation} -\dot M_2 = 2.64~10^{-8}\,{\rm M_{\odot}} {\rm yr^{-1}}\, \left(\frac{M_{\rm c,i}}{0.25{\rm M_{\odot}}}\right)^7\, \frac{M_1}{5M_1 - 6M_2}\, \left(\frac{M_2}{{\rm M_{\odot}}}\right)\, \left(\frac{M_2}{M_{\rm 2,i}}\right)^{-7/3}\, \left(\frac{M_1}{M_{\rm 1,i}}\right)^{-14/5} \label{7.1} \end{equation} for Pop. I, and \begin{equation} -\dot M_2 = 1.40~10^{-8}\,{\rm M_{\odot}} {\rm yr^{-1}}\, \left(\frac{M_{\rm c,i}}{0.25{\rm M_{\odot}}}\right)^4\, \frac{M_1}{5M_1 - 6M_2}\, \left(\frac{M_2}{{\rm M_{\odot}}}\right)\, \left(\frac{M_2}{M_{\rm 2,i}}\right)^{-20/9}\, \left(\frac{M_1}{M_{\rm 1,i}}\right)^{-8/3} \label{7.2} \end{equation} for Pop. II chemical composition. Results of detailed numerical computations by WRS are shown in their figs. 3 and 4. Comparing the results obtained from our analytical solution for the same parameters with those of WRS shows that the differences are small, too small in fact to be adequately shown on the scale of these figures. That the differences in $L(t)$ and $R(t)$ are small was to be expected based on the quality of the fits used (see Figs. 1 and 2). That the differences in $-\dot M_2$ are small too is shown in Figs. 3 and 4. In that context we should emphasize that our fits of the $L(M_{\rm c})$ and $R(M_{\rm c})$ relations of WRS with power laws are the simplest possible ones (with integer powers $\lambda$ and $\rho$) and that no attempt has been made to adjust the power law parameters such as to minimize the differences between the numerical results of WRS and the analytical solution. With $-\dot M_2$ being in close agreement with the numerical results it follows that also $M_2(t)$ and $P(t)$ should agree very well, which is indeed the case. Finally also $M_{\rm c}(t)$ must be in close agreement with the numerical result of WRS, otherwise the temporal evolution of the other quantities would not match either. \begin{figure \begin{minipage}[t]{0.48\hsize} \centerline{\epsfxsize=1.0\hsize\epsffile{Fig3.eps}} \caption{Comparison of the result obtained with the analytical solution (full line) with that of a matching numerical one by WRS (dashed line). Shown is the mass loss rate from a Pop. I giant as a function of the time for the parameters used by WRS in their figure 3, i.e. $X = 0.70$, $Z = 0.02$, $M_{\rm 1,i} = 1.4 {\rm M_{\odot}}$, $M_{\rm 2,i} = 1 {\rm M_{\odot}}$, and $M_{\rm c,i} = 0.26 {\rm M_{\odot}}$.} \end{minipage} \hfill \begin{minipage}[t]{0.48\hsize} \centerline{\epsfxsize=1.0\hsize\epsffile{Fig4.eps}} \caption{Comparison of the result obtained with the analytical solution (full line) with that of a matching numerical one by WRS (dashed line). Shown is the mass loss rate from a Pop. II giant as a function of the time for the parameters used by WRS in their figure 4, i.e. $X = 0.80$, $Z = 10^{-4}$, $M_{\rm 1,i} = 1.4 {\rm M_{\odot}}$, $M_{\rm 2,i} = 0.9 {\rm M_{\odot}}$, and $M_{\rm c,i} = 0.31 {\rm M_{\odot}}$.} \end{minipage} \end{figure} In fact we can summarize the result of this comparison as follows: if the power law fits to the core mass-luminosity and core mass-radius relations obtained from full stellar structure calculations are adequate, the analytical model provides an adequate approximation to the results of numerical computations. \section{SELFCONSISTENCY AND APPLICABILITY OF THE MODEL} \label{sect8} \subsection{Deviation from thermal equilibrium} \label{sect8.1} The equations we have derived in section \ref{sect5} are valid only to the extent that the donor star is not significantly driven out of thermal equilibrium as a consequence of mass loss. That is to say that the time-scale of mass loss \begin{equation} \tau_{\rm M_2} = -\frac{dt}{d \ln M_2} \label{8.1} \end{equation} must be significantly longer than the shortest time-scale on which the donor's convective envelope can react, i.e. longer than the Kelvin--Helmholtz time of the convective envelope (e.g. King et al. 1997a) \begin{equation} \tau_{\rm KH,ce} = \frac{3}{7}\, \frac{M_{\rm ce}}{M}\, \tau_{\rm KH} = \frac{3}{7}\, \frac{M - M_{\rm c}}{M}\, \frac{G M^2}{R L}\,. \label{8.2} \end{equation} Here $M_{\rm ce} = M - M_{\rm c}$ is the mass of the convective envelope and $\tau_{\rm KH}$ the Kelvin--Helmholtz time of the whole star. A measure for the star's deviation from thermal equilibrium is thus the the ratio $\tau_{\rm M_2}/\tau_{\rm KH,ce}$. Using equations (\ref{2.9}) and (\ref{3.6}) -- (\ref{3.8}) in (\ref{8.1}) and (\ref{8.2}) we can now write \begin{equation} \tau_{\rm M_2} = \left(\zeta_{\rm e} - \zeta_{\rm R,2}\right)\, \frac{X Q {\rm M_{\odot}}}{\rho L_0}\, \left(\frac{M_{\rm c,i}}{{\rm M_{\odot}}}\right)^{1-\lambda}\, \left(1 - \frac{t}{t_{\infty}}\right) \label{8.4} \end{equation} and \begin{equation} \tau_{\rm KH,ce} = \frac{3}{7}\, \frac{G {{\rm M_{\odot}}}^2}{R_0 L_0}\, \frac{M_2 - M_{\rm c}}{M_2}\, \left(\frac{M_2}{{\rm M_{\odot}}}\right)^{2-\zeta_{\rm e}}\, \left(\frac{M_{\rm c,i}}{{\rm M_{\odot}}}\right)^{-\left( \lambda + \rho \right)}\, \left(1 - \frac{t}{t_{\infty}}\right)^{\frac{\lambda + \rho}{\lambda -1}}\,, \label{8.5} \end{equation} and, therefore, \begin{equation} \frac{\tau_{\rm M_2}}{\tau_{\rm KH,ce}} = \left(\zeta_{\rm e}-\zeta_{\rm R,2}\right)\, \frac{7}{3}\, \frac{X Q R_0}{\rho G {\rm M_{\odot}}}\, \frac{M_2}{M_2 - M_{\rm c}}\, \left(\frac{M_2}{{\rm M_{\odot}}}\right)^{\zeta_{\rm e}-2}\, \left(\frac{M_{\rm c,i}}{{\rm M_{\odot}}}\right)^{1+\rho}\, \left(1 -\frac{t}{t_{\infty}}\right)^{\frac{1+\rho}{1-\lambda}}\,. \label{8.6} \end{equation} Inserting now for illustrative puposes the parameters of Table 2 which correspond to the results of WRS into (\ref{8.6}) yields \begin{equation} \frac{\tau_{\rm M_2}}{\tau_{\rm KH,ce}} = 3.16~10^3\, \left(-\zeta_{\rm R,2}\right)\, \frac{M_2}{M_2 - M_{\rm c}}\, \left(\frac{M_2}{{\rm M_{\odot}}}\right)^{-2}\, \left(\frac{M_{\rm c,i}}{0.25~{\rm M_{\odot}}}\right)^6\, \left(1 - \frac{t}{t_{\infty}}\right)^{-6/7} \label{8.7} \end{equation} for Pop. I, and \begin{equation} \frac{\tau_{\rm M_2}}{\tau_{\rm KH,ce}} = 3.83~10^3\, \left(-\zeta_{\rm R,2}\right)\, \frac{M_2}{M_2 - M_{\rm c}}\, \left(\frac{M_2}{{\rm M_{\odot}}}\right)^{-2}\, \left(\frac{M_{\rm c,i}}{0.25~{\rm M_{\odot}}}\right)^4\, \left(1 - \frac{t}{t_{\infty}}\right)^{-1} \label{8.8} \end{equation} for Pop. II chemical composition. What (\ref{8.6}) -- (\ref{8.8}) demonstrate is that for all practical purposes $\tau_{\rm M_2}/\tau_{\rm KH,ce} \gg 1$ unless the mass ratio is such that the binary system is close to thermal instability, i.e. that $\zeta_{\rm e} - \zeta_{\rm R,2} \ll 1$. Thus mass transfer driven by nuclear evolution of a star on the first giant branch is always sufficiently slow for the donor star to remain essentially in thermal equilibrium. To this extent, the analytical model presented in section \ref{sect5} is selfconsistent. \subsection{Irradiation of the donor star} \label{sect8.2} Since one of the main areas of application of this model is the evolution of luminous X-ray binaries (see e.g. Taam 1983 and WRS) and their transformation into millisecond pulsar binaries (see e.g. RPJDH and references therein) we have to examine whether this model remains applicable if irradiation of the cool donor star by X-rays from the vicinity of a very compact accretor is taken into account. As has first been shown by Podsiadlowski (1991) cool stars, when exposed to sufficiently strong external heating by X-rays, respond by expansion on the thermal time-scale of the convective envelope, thereby giving rise to very high mass transfer rates in a semi-detached binary. In the context of mass transfer from a giant to a compact accretor the stability of mass transfer with taking into account irradiation has been examined in some detail by King et al. (1997a). Their finding was that if the accretion rate on to the compact star equals the mass transfer rate at any time, the evolutionary model discussed here is not in general applicable, because mass transfer is very likely to be unstable. Rather than being continuous, mass transfer occurs in cycles which are characterized by short high states during which very high mass transfer rates are achieved and long low states during which the binary is detached. As has been shown by King et al. (1997a) this instability occurs mainly because for giants the ratio of mass loss time-scale to Kelvin--Helmholtz time (equation \ref{8.6}) is so large. It is therefore somewhat ironic that self-consistency of the model of nuclear time-scale driven stationary mass transfer from a giant is possible only at the price that such systems are in danger of being unstable to irradiation-driven mass transfer. However, the above-mentioned condition, namely that the accretion rate equals the mass transfer rate at any time, is not likely to be fulfilled. Rather, as has been shown by King et al. (1997b), the accretion disc in such binary systems is likely to be unstable and that even for the rather optimistic assumptions made by King et al. (1997b) about the efficiency of X-ray irradiation of the accretion disc. Based on newer results of Dubus et al. (1998) one could conclude that the accretion discs in essentially all systems with a giant donor are unstable. If this is the case, then accretion on to the compact star is spasmodic because the disc undergoes a dwarf nova-like limit-cycle instability. Estimates of the properties of the outbursts (e.g. King \& Ritter 1998) show that they are characterized by a very small duty cycle. This, in turn, is now of great importance for the stability of mass transfer from the giant in the presence of X-ray irradiation: King et al. (1997a) show that if accretion on to a neutron star or a black hole is spasmodic with a small duty cycle, then mass transfer from the giant is stable despite the donor being irradiated. If that is the case, the model considered by Taam (1993), WRS and us is applicable. On the other hand, if the accretor is a white dwarf the situation is more complicated: If the core mass of the donor is $M_{\rm c} \la 0.2 {\rm M_{\odot}}$, irradiation cannot destabilize mass transfer and, therefore, the model is applicable. Yet if $M_{\rm c} \ga 0.2 {\rm M_{\odot}}$, mass transfer is unstable. Although spasmodic accretion has a stabilizing effect, for mass transfer to become stable in this case the required small duty cycle has to be smaller still by a factor of $\sim 10 \cdots 100$ than in the case of a black hole or a neutron star (see King et al. 1997a, fig. 4). Because the value of the duty cycle does not depend explicitely on the nature of the accretor but rather on the average mass transfer rate and on the size of the accretion disc, it appears unlikely that the duty cycle in a white dwarf system is systematically smaller than in a black hole or neutron star system. Therefore it is more likely that in a white dwarf system in which $M_{\rm c} \ga 0.2 {\rm M_{\odot}}$ mass transfer is unstable despite spasmodic accretion. In this case the model in question is not applicable. However, we should emphasize that detailed numerical calculations of mass transfer from a giant with taking into account the effects of irradiation of the donor star and of disc instabilites have yet to be done. Before results of such computations are available, a final decision about whether this model is applicable cannot be made. \section{DISCUSSION} \label{sect9} Having established that the analytical approximation is both accurate and selfconsistent, and that the model is probably also applicable to the most important case of interest, i.e. to low-mass X-ray binaries, we can now proceed to discuss the properties of these solutions. One of the remarkable properties of this type of evolution which has already been pointed out by WRS is that the mass transfer rate remains almost constant for most of the mass transfer phase. Here we shall examine whether this is a generic property of mass transfer from a giant or whether this is due to a particular choice of parameters. Other aspects which we wish examine are the dependence of the temporal evolution on the initial parameters, in particular on the initial core mass $M_{\rm c,i}$, and systematic differences between Pop. I and Pop. II systems. \begin{figure \centerline{\epsfxsize=0.5\hsize\epsffile{Fig5.eps}} \caption{Mass transfer rate computed from the analytical solution for three different cases: The full and dashed lines show respectively the mass transfer rate as a function of time in an evolution of a Pop. I and a Pop. II system with same initial parameters as those used in Figs. 3 and 4, but now for non-conservative mass transfer characterized by $\eta = 0$ and $\nu = 1/q$. The dash-dotted line shows the mass transfer rate in a conservative evolution of a Pop. I system with the parameters $M_{\rm 1,i} = 5 {\rm M_{\odot}}$, $M_{\rm 2,i} = 2 {\rm M_{\odot}}$, and $M_{\rm c,i} = 0.26 {\rm M_{\odot}}$.} \end{figure} \subsection{The mass transfer rate as a function of time} \label{sect9.1} One of the remarkable properties of the evolutions shown by WRS in their figs. 3 and 4 (see also our Figs. 3 and 4) is that the mass transfer rate remains constant within a small factor ($\la 2$) in the case of conservative mass transfer, and, as Fig. 5 shows, within an even smaller factor ($\la 1.5$) in the case that $\eta = 0$ and $\nu = 1/q$ throughout the evolution. This raises the question why this is so and whether this is a generic property of this type of evolution. For our discussion we combine (\ref{2.9}) and (\ref{3.8}) to write the mass transfer rate as \begin{equation} -\dot M_2 = \frac{M_2}{\zeta_{\rm e} - \zeta_{\rm R,2}}\, \frac{\rho}{\lambda -1}\, \frac{1}{t_{\infty} - t}\,. \label{9.1} \end{equation} From (\ref{9.1}) it is immediately seen, that it contains two counteracting factors: first, with time $\zeta_{\rm e} - \zeta_{\rm R,2}$ increases because q increases. This factor has thus the tendency to reduce $-\dot M_2$ with time. On the other hand, $t_{\infty} - t$ decreases with time and thus has the tendency to increase $-\dot M_2$. Therefore, there is a possibility that $-\dot M_2$ goes through a minimum during an evolution. Yet, an evolution does not necessarily include this minimum. Rather, whether or not $-\dot M_2(t)$ has a minimum depends on the initial mass ratio $q_i$ and on the subsequent change of $q$ with time. Starting from equations (\ref{5.2}) and (\ref{5.3}) and working out $\ddot M_2 = 0$ leads to a (rather complicated) equation for the mass ratio $q_0$ at which $-\dot M_2$ becomes extremal. From the above discussion we know already that this extremum is a minimum. Denoting the final mass ratio by $q_{\rm f} = M_{\rm 1,f}/M_{\rm 2,f} \approx M_{\rm 1,f}/ M_{\rm c,f}$, we can now distinguish three different cases: 1) if $q_{\rm crit} < q_{\rm i} < q_0 < q_{\rm f}$ then the mass transfer rate goes through a minimum, 2) if $q_{\rm crit} < q_0 < q_{\rm i} < q_{\rm f}$, the mass transfer rate will increase monotonically, and if 3) $q_{\rm crit} < q_{\rm i} < q_{\rm f} < q_0$, the mass transfer rate will decrease monotonically. Here $q_{\rm crit}$ is the critical mass ratio for both thermal and adiabatic stability against mass transfer. Its value is given by the larger of the positive solutions of the equations $\zeta_{\rm ad} - \zeta_{\rm R,2} = 0$ and $\zeta_{\rm e} - \zeta_{\rm R,2} = 0$ (see e.g. Ritter 1988). For illustrative purposes we give the values of $q_{\rm crit}$ and $q_0$ for the cases we show in Figs. 3 -- 5. For these cases we have assumed with WRS $\zeta_{\rm e} = 0$ and implicitly $\zeta_{\rm e} < \zeta_{\rm ad}$. From this and (\ref{4.12}) it follows that $q_{\rm crit} = 1.2$ in the case of conservative mass transfer, and with (\ref{4.14}) that $q_{\rm crit} = 0.836$ in the case where $\eta = 0$ and $\nu = 1/q$. The corresponding values of $q_0$ are in the case of conservative mass transfer $q_0 = 3.77$ for Pop. I giants (with parameters taken from Table 2) and $q_0 = 3.95$ for Pop. II giants. In fact for all combinations $(\lambda , \rho)$ which describe the evolution of giants adequately we find $q_0 \la 4$. In the case of non-conservative mass transfer with $\eta = 0$ and $\nu = 1/q$ we find $q_0 = 1.89$ for Pop. I giants and $q_0 = 1.97$ for Pop. II giants, again with parameters taken from Table 2, and an upper limit of $q_0 \la 2$ for all acceptable combinations of $(\lambda , \rho)$. From the foregoing it is now clear that the near constancy of the mass transfer rate in the evolutions considered by WRS is not a generic property. Rather it is the result of a particular choice of initial parameters. In particular, if $q_{\rm i}$ is close to (but larger than) $q_{\rm crit}$ the mass transfer rate initially drops very strongly. On the other hand, the larger $q_{\rm f}$ the more strongly the mass transfer rate increases beyond $q_0$. Because conservative mass transfer results in the largest possible values of $q_{\rm f}$ we expect the mass transfer rate to increase substantially beyond $q_0$ and that in particular if the initial mass of the primary is already much larger than that of the donor star. This is e.g. the case if the accretor is a black hole rather than a neutron star. To illustrate this we show in Fig. 5 the run of $-\dot M_2$ in a conservative mass transfer with $M_{\rm 1,i} = 5{\rm M_{\odot}}$, $M_{\rm 2,i} = 2{\rm M_{\odot}}$ and $M_{\rm c,i} = 0.26{\rm M_{\odot}}$. It is clearly seen how the mass transfer rate grows because of the high final mass ratio of $q_{\rm f} = 25.9$. Nevertheless, the mass transfer rate stays reasonably constant if $q_{\rm i}$ is not close to $q_{\rm crit}$ and the accretor is a neutron star, i.e. $M_{\rm 1,i} \approx 1.4 {\rm M_{\odot}}$. This is the case for the simulations shown by WRS. On the other hand, in the case that $\eta = 0$ and $\nu = 1/q$ the variation of the mass transfer rate is systematically smaller, first because for the same $q_{\rm i}$ mass transfer is more stable than in the conservative case, and second because $M_1$ remains constant, thus resulting in systematically lower values of $q_{\rm f}$. This is illustrated in Fig. 5 where we show the run of $-\dot M_2$ for the same initial parameters as for the conservative evolutions shown in Figs. 3 and 4, but now with $\eta = 0$ and $\nu = 1/q$. As can be seen, the mass transfer varies much less than in the conservative case. \subsection{Dependence on the initial core mass} \label{sect9.2} As can be seen from equation (\ref{5.3}), apart from the variation of the mass transfer rate during a mass transfer phase discussed in section \ref{sect9.1}, the level of mass transfer is set by the initial core mass $M_{\rm c,i}$ and depends on a large power, i.e. $\lambda -1$, of it. Using the parameters given in Table 2 which correspond to the results of WRS we have respectively $-\dot M_2 \propto {M_{\rm c,i}}^7$ and $-\dot M_2 \propto {M_{\rm c,i}}^4$ for Pop. I and Pop. II giants (see equations (\ref{7.1}) and (\ref{7.2})). This strong dependence on $M_{\rm c,i}$ is a direct consequence of the strong dependence of the stellar luminosity on $M_{\rm c,i}$ (see equation (\ref{3.1})). Also the stellar radius shows a strong dependence on $M_{\rm c,i}$ as can be seen from equation (\ref{3.7}). Therefore, the initial core mass is the single most important parameter characterizing the evolution of a binary with mass transfer from a giant. \subsection{Dependence on $\zeta_{\rm e}$} \label{sect9.3} We have already discussed the possible influence of $\zeta_{\rm e}$ on the value of $q_{\rm crit}$ in section \ref{sect9.1}. However, as Table 1 shows, the analytical solutions (\ref{5.2}) -- (\ref{5.6}) depend on $\zeta_{\rm e}$ also via the exponent $p_2$. Going from a solution with $\zeta_{\rm e} = 0$ to one with $\zeta_{\rm e} \ne 0$ changes $p_2$ by $\Delta p_2 = - \zeta_{\rm e} (\lambda -1)/\rho$. Thus, with the parameters from Table 2 which correspond to the results of WRS we have $\Delta p_2 = -{7/5}\,\zeta_{\rm e}$ for Pop. I, and $\Delta p_2 = -{4/3}\, \zeta_{\rm e}$ for Pop. II chemical composition. Taking for an estimate $\zeta_{\rm e} \approx -0.3$ we find that for both cases, i.e. Pop. I and Pop. II, $\Delta p_2 \approx 0.4$. Therefore, apart from increasing $q_{\rm crit}$, going from $\zeta_{\rm e} = 0$ to $\zeta_{\rm e} < 0$ leads to an increase of the mass transfer rate via the increased factor $1/(\zeta_{\rm e} - \zeta_{\rm R,2})$ and to a decrease by the factor $(M_2/M_{\rm 2,i})^{\Delta p_2}$. Sufficiently far from the stability limit, i.e. from $q_{\rm crit}$, it is the latter factor which dominates. Because $\zeta_{\rm e} \ga -0.3$ this factor is never very small. For all practical purposes we find it to be between $\sim 0.5$ and $\sim 0.7$. Thus, qualitatively, the main effect of going from $\zeta_{\rm e} = 0$ to $\zeta_{\rm e} < 0$ is to increase $q_{\rm crit}$ and therefore to reduce the maximum allowed value for the initial secondary mass. \subsection{Pop. I versus Pop. II systems} \label{sect9.4} Because for Pop. II giants $L_0$, $R_0$, $\lambda$ and $\rho$ are significantly smaller than for Pop. I giants (see Table 2), the evolution of a Pop. II system, all other parameters being the same, is significantly slower than that of a Pop. I system if $M_{\rm c,i} \ga 0.2 {\rm M_{\odot}}$. From equation (\ref{5.3}) it follows that the differential evolutionary speed scales as ${M_{\rm c,i}}^{\lambda_{II} -\lambda_I}$, i.e. using the parameters of Table 2 which correspond to the results of WRS, as ${M_{\rm c,i}}^{-3}$. For an initial core mass of $0.3 {\rm M_{\odot}}$ and $0.4 {\rm M_{\odot}}$ the evolution of a Pop. II system is thus slower by respectively a factor of about 3 and 8. Furthermore, because Pop. II giants are also smaller than Pop. I giants if $M_{\rm c} \ga 0.2 {\rm M_{\odot}}$ the orbital period of a Pop. II binary is shorter too by a corresponding factor of about $(M_{\rm c,i}/{0.2 {\rm M_{\odot}}})^{-3}$ than that of a comparable Pop. I system. \section{CONCLUSIONS} \label{sect10} We have shown that the evolution of a binary with stable mass transfer from a giant donor can be described by a simple analytical solution (see section \ref{sect5}) if the core mass-luminosity and core mass-radius relation are simple power laws of the mass of the degenerate helium core $M_{\rm c}$. Furthermore, we have seen that the results of full stellar structure calculations of giants given by WRS and others can indeed be approximated with little loss of accuracy by simple power laws (section \ref{sect6}) and that, therefore, the analytical solutions derived here are relevant for the cases studied by Taam (1983) and WRS and similar cases. Moreover, the analytical solutions which we have derived are more general than the numerical ones of WRS in the sense that we allow for non-conservative mass transfer and a non-zero mass radius exponent $\zeta_{\rm e}$. A direct comparison of the results obtained from the analytical solution with those of numerical calculations by WRS (section \ref{sect7}) shows that both the analytical solution and the power law approximations are quite accurate (see Figs. 1 -- 4). Using the analytical solution we have also made an a posteriori check of whether one of the key assumptions made in this model is fulfilled, namely that the donor star remains near thermal equilibrium despite being subject to mass loss (section \ref{sect8.1}), and have found that this is indeed the case. In section \ref{sect8.2} we put forward arguments, why, despite strong irradiation of the donor star from a compact accretor, the model is probably applicable to low-mass X-ray binaries. Furthermore we have seen (section \ref{sect9.1}) that the near constancy of the mass transfer rate over most of the mass transfer phase seen in the results by WRS is not a generic feature of this type of evolution but rather a consequence of a particular choice of parameters. In particular, we found considerable variation of the mass transfer rate if the final mass ratio $q_{\rm f}$ gets large. Therefore, particularly in conservative evolutions in which the accretor is massive, e.g. a black hole, the mass transfer rate increases quite substantially towards the end of the mass transfer phase (see Fig. 5). On the other hand, in evolutions in which the mass of the accretor remains constant or nearly so, $q_{\rm f}$ gets much less extreme and as a consequence of that the mass transfer rate varies to a much lesser degree. From the analytical solution (\ref{5.3}) it is also directly seen that the level of mass transfer is largely set by $M_{\rm c,i}$, the core mass at the {\em onset} of mass transfer. \section{Acknowledgments} The author thanks Andrew King, Uli Kolb and Hans--Christoph Thomas for helpful discussions, the Leicester University Astronomy Group for its hospitality during 1998 November/December when part of this work was done, and for support from its PPARC Short-Term Visitor grant.
\section{Introduction} Although the translational symmetry of a crystalline solid imposes a delocalized basis of Hamiltonian eigenstates (Bloch's functions), it is sometimes advantageous to consider a transformation to a new set of basis functions with a local character. Beyond the mathematical equivalence (both sets span the same space of states), a local viewpoint is better suited for the analysis of concepts such as bonding which are eminently local in character. Recent work on electronic Wannier functions has shown the usefulness of a local representation in the chemical characterization of a given band subspace,~\cite{marzari} in the analysis of bonding topology in a disordered system,~\cite{parrinello} and in more formal developments.~\cite{other-developments} The lattice dynamical problem is formally very similar to the electronic one: a set of Bloch eigenstates (normal modes) represents the collective vibrations of the atoms in the crystal. A basis change to a set of local displacement patterns (lattice Wannier functions or local modes) can in principle be achieved. So far, the main application for these local modes has been in the field of structural phase transitions. Typically, the behavior of a given dispersion branch or set of branches determines the essential instabilities of the system, and the associated degrees of freedom enter into the construction of an effective Hamiltonian which reproduces the relevant physics. Through the use of a localized basis set, the number of coupling terms in the effective Hamiltonian can be relatively small, easing the statistical mechanical treatment and the interpretation of the results. In particular, the anharmonic terms in the effective Hamiltonian can be kept local (on-site), in contrast with what happens in a reciprocal-space description. This local mode approach has been used extensively in the past to gain an understanding of the behavior of complex systems, but until recently the local variables were treated as dummy degrees of freedom in a semi-empirical model, with their interactions fitted to reproduce the observed phenomena. In the last few years, a new approach, in which the effective Hamiltonian is parametrized on the basis of first principles calculations, has had great success in studies of the phase transition sequences in perovskite oxides.~\cite{zhong_batio3,rabe_waghmare,zhong_srtio3} Central to the parametrization process is the explicit construction of lattice Wannier functions, and two schemes have been proposed to carry it out. Zhong, Vanderbilt, and Rabe~\cite{zhong_batio3} (ZVR) used the structure of the zone-center soft mode in perovskite BaTiO$_3$ to construct symmetry-adapted highly localized local modes. Subsequently, Rabe and Waghmare~\cite{rabe_wannier} (RW) generalized this approach to reproduce the normal modes at several (typically, high symmetry) points of the Brillouin Zone. While both the ZVR and RW approaches have been broadly successful in the specific problems for which they were conceived, in this paper we will argue that they are not completely satisfying in some respects. We will present a new procedure to generate lattice Wannier functions, an approach which makes use of the available symmetry information, produces local modes with a high degree of localization, enables a systematic improvement of their quality, and is straightforward to implement. \section{Method} We are interested in describing a {\sl relevant subspace} $\cal R$ of the full $3Np$-dimensional configuration space of a crystal with $p$ atoms per unit cell. Typically, we can choose ${\cal R}$ as a complete {\sl band} of dispersion branches (complete in the sense that it is invariant under the action of the space group of the crystal~\cite{explain-complete}). Associated to a branch $j$ is a set of {\sl normal modes} ($3Np$-dimensional vectors) $\{{u}_j^{\bf k}\}$~\cite{vector-notation} which are eigenvectors of the Fourier transform of the force-constant matrix.~\cite{force-constant} (The displacement in the $\alpha$ cartesian direction of the atom $\kappa$ in cell ${\bf l}$ is given explicitly by ${u}_j^{\bf k}({\bf l},\kappa,\alpha)$.) The normal modes transform according to irreducible representations of the little groups $G^{\bf k}$. These representations, considered over the whole BZ, determine the {\sl band symmetry}. The relevant subspace ${\cal R}$ is spanned by all the $\{{u}_j^{\bf k}\}$ in the band, but it is clear that any transformation \begin{equation} \tilde u_j^{\bf k} = \sum_{i=1}^n{M_{ji}^{\bf k} {u}_i^{\bf k}} \label{eq:bloch_modes} \end{equation} will lead to a new basis of extended states which we will call {\sl Bloch modes}. Here $n$ is the {\sl band dimension}, the number of dispersion branches in the band. Having thus specified the relevant subspace by means of Fourier space variables $\{u_j^{\bf k}\}$, the problem we tackle is the construction of a new basis $\{w_j^{\bf n}\}$ which is local, as opposed to extended, in character. Mathematically, the $\bf k$ label should be exchanged by a local label ${\bf n}$ associated to the different unit cells in the crystal. Translational symmetry takes the form: \begin{equation} {w}_j^{\bf n}({\bf l},\kappa,\alpha) = {w}_j^{{\bf n}+{\bf t}}({\bf l}+{\bf t},\kappa,\alpha) \; , \label{eq:w_trans} \end{equation} which is trivially satisfied by the standard Wannier function form~\cite{wannier}: \begin{equation} {w}_j^{\bf n} = {1\over\Omega}\int_{BZ} {\exp(-i{\bf kn})\tilde{u}_j^{\bf k}d{\bf k}} \label{eq:w_def} \end{equation} in which $\Omega$ is the volume of the BZ. A high degree of localization means that the displacement ${w}_j^{\bf n}({\bf l},\kappa,\alpha)$ should be very small or zero when ${\bf l}$ is a few lattice constants away from ${\bf n}$. The arbitrariness implicit in their definition (Eq.~\ref{eq:bloch_modes}) means that the Wannier functions are non-unique, and a relatively large latitude then exists to tune their properties. In particular, the degree of localization has traditionally been the focus of great interest, and recently, Marzari and Vanderbilt~\cite{marzari} have succeeded in optimizing the matrices appearing in Eq.~\ref{eq:bloch_modes} to construct very localized electronic Wannier functions starting from the Bloch states. A restriction to unitary matrices resulted in an orthonormal basis of Wannier functions, and the optimization process led to symmetric-looking functions, even though no symmetry conditions were explicitly imposed.~\cite{dhv_start} In principle, such an approach should work for the vibrational problem, too. However, we prefer to take an alternate route which takes advantage of the knowledge of the band symmetry. \subsection{Symmetry requirements} As studied extensively in the literature,~\cite{zak} one should supplement the translational constraints of Eq.~\ref{eq:w_trans} with another set of conditions which represent the transformational properties of the ${w}_j^{\bf n}$ under the effect of the point symmetry of the crystal. These are most easily discussed by introducing a symmetry-based definition of the {\sl center} of a mode. Consider a Wyckoff set with representative site $\bf r$ and the set $\hat{G}_{\bf r}$ of operations in $G$ that leave $\bf r$ invariant.~\cite{site-symmetry-group} Given an irreducible representation $\tau$ of $\hat{G}_{\bf r}$ with dimension $d_{\tau}$, any $d_{\tau}$ displacement patterns transforming with $\tau$ under the action of $\hat{G}_{\bf r}$ are said to be centered in $\bf r$. It is then notationally more convenient to use a double index to label these patterns: ${w}_{{\bf r}, s}$ where $s$ ranges from 1 to $d_{\tau}$. The action of the elements of the space group $G$ on this set generates images at the rest of the positions in the Wyckoff set, i.e., $d_rd_{\tau}$ patterns $w^{\bf n}_{{\bf r}_i,s}$ per cell, where $i$ ranges from 1 to $d_r$ (the multiplicity of the Wyckoff set). This set of lattice functions is represented by the pair $(\bf r,\,\tau)$ and define a representation of $G$ which is called {\sl band representation}.~\cite{zak} A necessary condition for the description of a relevant band subspace by means of these symmetry-adapted local modes is the equivalence of the band symmetry of $\cal R$ and the band representation $(\bf r,\,\tau)$ (in particular this implies $n\,=\,d_rd_{\tau}$). More details about the choice of the correct $(\bf r,\,\tau)$ for a given $\cal R$ are presented in the Appendix, where we also discuss the transformation properties of the corresponding $\{\tilde u_j^{\bf k}\}$. Incidentally, since Eq.~\ref{eq:w_def} establishes a correspondence between lattice Wannier functions and Bloch modes, in what follows the latter can be also labeled by the site and representation indexes: $\tilde{u}^{\bf k}_{{\bf r},s}$ \subsection{Practical criterion for localization} A straightforward scheme to obtain lattice Wannier functions can be based on a direct use of Eq.~\ref{eq:w_def}, performing the BZ sum by means of any of the standard ``special k-points'' methods.~\cite{chadi-cohen,monkhorst-pack} The quality of the subspace description can thus be systematically improved by simply using denser k-point sets. This approach can incorporate information about the normal modes throughout the whole Brillouin zone, as opposed to at just one point (as in the ZVR method), or at a very special set of high-symmetry k-points (as in the RW scheme). As stated in the Introduction, it is highly desirable that the local mode basis functions be as localized as possible, in order to permit the consideration of only a few coupling terms in the effective Hamiltonian. From the point of view of real applications, a basis of Wannier functions which are not localized is not efficient, even if it spans $\cal R$ perfectly. The form of Eq.~\ref{eq:w_def} suggests a very simple heuristic criterion to achieve a high degree of localization for the lattice Wannier functions: choose the $M^{\bf k}$ matrices in such a way that the $\tilde{u}^{\bf k}$ Bloch vectors at different ${\bf k}$ add their contributions coherently at the center of the Wannier function. Interference effects can then be counted on to automatically dampen the amplitude of the displacements at sites away from the center. Both the symmetry requirements and the localization condition can be formulated in the following way. Assume a $({\bf r}, \tau)$ pair has been determined on the basis of band symmetry, and that we focus on the construction of local modes at cell $\bf n$=$\bf 0$. Consider a set of $d_{\tau}$ ($3Np$-dimensional) orthonormal vectors $\{ x_{{\bf r},s}\}$ which are centered in $\bf r$, transform with irrep $\tau$ and involve atoms in an orbit as close to $\bf r$ as possible.~\cite{x-orbit} The localization criterion is implemented by requiring that $\tilde{u}^{\bf k}_{{\bf r},s}$ be orthogonal to $x_{{\bf r},t}$ if $s\neq t$, and ``parallel'' (meaning that their scalar product is positive) if $s=t$.~\cite{x-criterion} It can be seen that this condition fixes the form of the $M^{\bf k}$ matrices, up to overall normalization factors. In general, the $M^{\bf k}$ will not be unitary, with the result that two lattice Wannier functions at different cells $w_{{\bf r},s}^{\bf n}$ and ${w}_{{\bf r}',s'}^{\bf n'}$ will not be orthogonal if the pairs $({\bf r},s)$ and $({\bf r}',s')$ are not equal.~\cite{non-orthog} In the next section we will provide a simple worked example of the new construction scheme and will compare its results to those of other methods. \section{Examples and Discussion} In order to illustrate the scheme presented in the previous section, we will employ a two-dimensional model crystal with two different atoms which occupy the $1a$ (white) and $1b$ (black) Wyckoff positions of the plane group $p4mm$ (See Fig.~\ref{fig:crystal} a)). A simple harmonic model for the force constants (with the couplings among the white atoms considered up to fourth nearest neighbors and the rest to first nearest neighbors, which corresponds to 6 independent parameters) gives the dispersion branches of panel b) in the figure. We will focus our attention on the two optical branches, which form a single band since they are essentially degenerate at the $\Gamma$ and M points. These optical branches transform according to the decompositions \begin{equation} \begin{array}{lll} \Gamma&(4mm):& E\cr {\rm X}&(2mm):&B_1+B_2\cr {\rm M}&(4mm):&E \;, \end{array} \label{eq:bz_map} \end{equation} in irreducible representations of the little co-groups at the high symmetry points. A simple application of the procedure spelled out in the Appendix shows that the band representation compatible with the above band symmetry is that represented by the pair $(\circ,E)$, in which $E$ is a two-dimensional irreducible representation which turns out to be the vector representation of the point symmetry group at $\circ$. The set of $\{x\}$ vectors is then trivial to construct: as the ``$\circ$'' Wyckoff position is occupied, it is just enough to make $x_{{\circ},1}$ and $x_{{\circ},2}$ unit vectors attached to the central atom and pointing in the $x$ and $y$ cartesian directions, respectively. For this crystal structure, the simplest non-trivial set of special k-points is given by $\{(1/8,1/8);(1/8,3/8);(3/8,3/8)\}$. The explicit application of the localization criterion proceeds as follows. At each k-point in the set the normal modes are computed and the $M^{\bf k}$ matrices constructed. For example, at $(1/8,3/8)$, the normal modes are \begin{equation} \begin{array}{ll} {u}^{\bf k}_{1} = &(\ldots;0.23,0.93;\ldots)\cr {u}^{\bf k}_{2} = &(\ldots;0.84,-0.19;\ldots)\; , \end{array} \label{eq:ex_modes} \end{equation} where the ``$\ldots$'' refer to displacements on atoms other than the one at the center. The ``coherent addition at the center'' condition then becomes: \begin{equation} \begin{array}{l} 0.23\,M_{11}+0.84\,M_{12} > 0\cr 0.23\,M_{21}+0.84\,M_{22} = 0\cr 0.93\,M_{11}-0.19\,M_{12} = 0\cr 0.93\,M_{21}-0.19\,M_{22} > 0\; , \end{array} \label{eq:m_cond} \end{equation} and is satisfied by \begin{equation} M=\left(\matrix{0.200&0.980\cr 0.964&-0.264\cr}\right) \; , \label{eq:m_matrix} \end{equation} uniquely defined but for row-specific arbitrary factors. Since M is not unitary, the two optical Bloch vectors $\tilde u_{\circ,s}$ at this k-point will not be orthogonal (although they can of course still be chosen to be normalized). Once this procedure has been performed at every k-point in the set, the integral (sum) in Eq.~\ref{eq:w_def} can be carried out to give the components of the lattice Wannier functions. Since the $\{\tilde u^{\bf k}\}$ determined by the localization criterion also satisfy the symmetry compatibility relations (see Appendix), the Wannier functions are symmetry-adapted. In Fig.~\ref{fig:shells}~c) we show the displacements associated to the local mode $w_{\circ,1}$, which transform as the first component of the vector representation $(E)$ of $4mm$. The degree of localization of these lattice Wannier functions can be gauged by computing the contribution to the total norm from a given shell around the center atom, as presented on Table~\ref{tab:shells}. Less than one per cent of the norm is outside the fourth shell (which corresponds roughly to the second-neighbor unit cells). If the integral in Eq.~\ref{eq:w_def} is computed using a denser special-point set, the degree of localization is maintained, as can be seen by comparing the columns labeled ``this work (3 k)'' and ``this work (10 k)'' on Table~\ref{tab:shells}. This means that the quality of the local modes can be systematically improved while retaining a high degree of localization. It is enlightening to compare this scheme to that of Rabe and Waghmare.~\cite{rabe_wannier} In the latter the analysis of the symmetry compatibility relations proceeds in the same way, and once the right $({\bf r},\tau)$ set has been identified, a series of orthonormal ${x}$ sets is constructed at successive shells centered on $\bf r$. The extent of the outermost shell fixes the localization of the Wannier functions by construction, and the actual atomic displacements are determined by fitting to the normal modes computed at a few high-symmetry points of the Brillouin Zone. In essence, the normal-mode information determines the weight assigned to each symmetry-adapted shell, so there is a tradeoff between the extent of the lattice Wannier functions and the amount of information from the real dispersion relations that can be used in the construction procedure. For example, in the PbTiO$_3$ work, Rabe and Waghmare found that adding information about the normal modes at the X point resulted in a less localized local mode than if only the $(\Gamma,{\rm M},{\rm R})$ set was used.~\cite{rabe_waghmare} In contrast, our scheme can deal with the extra k-point without loss of localization: our local modes for PbTiO$_3$ using four high-symmetry points~\cite{mp-unshifted} are more localized than the best (three point) RW lattice Wannier functions. It is clear that localization cannot be the main quality criterion for the construction of local modes. If it were, then the ZVR scheme, which uses only one high-symmetry k-point to construct a (very localized) lattice Wannier function, would be the method of choice. In fact, the real test for local mode sets is the degree to which they reproduce the energetics of the relevant subspace $\cal R$. That is, in our case, the degree to which the dispersion relations of the effective Hamiltonian \begin{equation} H_{\rm eff}=H_{\rm eff}(Q_1,Q_2,\ldots,Q_{Nn}) \label{eq:heff} \end{equation} match the real dispersion branches associated with $\cal R$. In Eq.~\ref{eq:heff}, the variables $Q_i$ are the amplitudes of the local mode variables, so that $H_{\rm eff}$ can be thought of as the ``projection'' of the complete Hamiltonian into the relevant subspace $\cal R$ (which is typically considered as energetically decoupled from the rest of the configuration space of the crystal). The explicit form of $H_{\rm eff}$ will depend on the detailed structure of the lattice Wannier functions. In particular, the number of distinct coupling coefficients (representing the interaction of modes at different sites) which one should take into account in $H_{\rm eff}$ is determined by the spatial extent of the local modes. We have constructed effective Hamiltonians for the model crystal for each of the three local-mode construction schemes discussed above (we obtain the coupling between $w^{\bf n}_{\circ,s}$ and $w^{{\bf n}'}_{\circ,s'}$ by calculating the energy associated to the crystal when it is distorted by just these modes). The original crystal Hamiltonian involved interactions up to fourth nearest neighbors for white atoms. Since the local modes involve basically displacements of the central white atom, we have kept the same range of interaction in $H_{\rm eff}$, but now referring of course to fourth nearest local modes. This amounts to using ten independent coupling coefficients; a larger number of parameters would not be reasonable in a practical application. ZVR-style local modes are very localized and do not couple beyond the fourth neighbor shell, so the considered $H_{\rm eff}$ includes all the existing interactions. This can be seen on panel a) of Figure~\ref{fig:gm}: the dispersion branches computed from $H_{\rm eff}$ match the exact ones at the $\Gamma$ point. However, $H_{\rm eff}$ gives a poor description of the dispersion branches away from $\Gamma$, as it should be expected in view of the construction procedure. (Incidentally, the inverse of Eq.~\ref{eq:w_def} leads to Bloch modes which are not normalized to unity, except at the $\Gamma$ point. The standard analysis of $H_{\rm eff}$ as given would lead to the low-lying dispersion branches in the figure. The higher branches are obtained by considering the corresponding generalized eigenvalue problem.) Panel b) shows that the $H_{\rm eff}$ constructed on the basis of RW local modes gives a good qualitative overall description of the dispersion, but fails to match the exact branches at the $\Gamma$ point (as it should, given that this point was used in the construction scheme). The reason is that the local modes are more extended, and it is necessary to include couplings to further shells (at least up to seventh nearest neighbors)~\cite{more-info} for the match to be essentially perfect. This means that the RW scheme does not lead to efficient local modes, in the sense stated above.~\cite{rabe-fit} The situation gets worse if more accuracy is needed in the overall description of the dispersion branches: the local modes turn out to be more extended, and even more coupling terms are needed in $H_{\rm eff}$. In contrast, the local modes constructed following our heuristic criterion for localization do exhibit good efficiency (the dispersion branches do not change much when couplings to more than fourth nearest neighbors are included) and provide a very good qualitative match of the true branches throughout the BZ (Fig.~\ref{fig:gm} c)). (It should be noted that our construction scheme does not involve any high-symmetry points, hence the offset of the branches at $\Gamma$ and M. We trade an overall good match for perfect accuracy at a few points.) Since a few coupling terms are enough to take into account the structure of the local modes, and the resulting $H_{\rm eff}$ provides a good fit to the true branches, our lattice Wannier functions are well suited for the local representation of the relevant subspace $\cal R$. Moreover, they can be improved if needed by including more k-points in the integration set, with only a minor sacrifice in the compactness of the effective Hamiltonian. We find these general conclusions to remain valid when more complicated interaction models are considered. The practical application of our method of local mode construction to real materials requires the knowledge of the normal modes at general points of the Brillouin Zone. This information is easily obtained with modern linear-response codes without the need for large supercells. In the field of phase transitions, the use of this new scheme should enable the study of more complicated situations than those considered up to now. Competition of instabilities associated to different regions of the BZ or complications derived from anti-crossing phenomena are examples in which this method is bound to be useful. On the other hand, this work might provide an illustration of some of its theoretical underpinnings: the physical interpretation of the band representation associated to a dispersion band~\cite{band-interpretation} or the symmetry-induced continuity of phonon spectra~\cite{michel-walker-zak} are two instances of this. \section{Conclusions} We have presented a straightforward scheme for the construction of very localized lattice Wannier functions, with explicit consideration of crystal symmetry. The new localization procedure enables a systematic improvement in the description of the relevant physics (by simply using denser sets of special k-points in the BZ integration) while still being quite efficient in regard to the number of coupling parameters needed in the effective Hamiltonian. Besides, the present method is straightforward to implement. \section*{Acknowledgements} We thank Karin Rabe, Philippe Ghosez, and David Vanderbilt for useful comments. This work was supported in part by the UPV research grant 060.310-EA149/95 and by the Spanish Ministry of Education grant PB97-0598. J.I. acknowledges fellowship support from the Basque regional government and thanks Agustin V\'algoma for comments on the manuscript. \section*{Appendix} Let us study in more detail the equivalence between the {\sl band symmetry} emerging from the transformation properties of the Bloch modes and the {\sl band representation} associated to a $({\bf r},\tau)$ set.~\cite{zak} This can be done by considering the action of $G$ on the $({\bf r},\tau)$ set of local modes and, consequently, on the associated Bloch modes $\tilde{u}^{\bf k}_{{\bf r}_i,s}$. In order to proceed, we need to formulate the transformation properties of the modes ${w}^{\bf 0}_{{\bf r},s}$ under the action of $\{\bar{R}|\bar{\bf v}\}\in \hat{G}_r$. Since we consider $\{\bar{R}|\bar{\bf v}\}$ acting on the modes themselves and not on their components, we denote symmetry operations by the associated operators $O\{\bar{R}|\bar{\bf v}\}$. We have \begin{equation} O\{\bar{R}|\bar{\bf v}\}\,{w}^{\bf 0}_{{\bf r},s}\; =\; \sum_{h=1}^{d_{\tau}}\,D^{\tau}_{hs}(\{\bar{R}|\bar{\bf v}\}) \,{w}^{\bf 0}_{{\bf r},h} \label{local-transformation} \end{equation} where ${\bf D}^{\tau}(\{\bar{R}|\bar{\bf v}\})$ is the matrix associated to $\{\bar{R}|\bar{\bf v}\}$ by irrep $\tau$. Now, we consider the rest of elements in the $({\bf r},\tau)$ set. They can be mathematically defined as \begin{equation} {w}^{\bf n}_{{\bf r}_i,s}\; :=\; O\{E|{\bf n}\}\,O\{R_i|{\bf v}_i\}\,{w}^{\bf 0}_{{\bf r},s} \end{equation} where $\{E|{\bf n}\}$ is a lattice translation and $\{R_i|{\bf v}_i\}$ is one of the $d_r$ elements in $G/G_r$, which are chosen so that all the ${\bf r}_i:=\{R_i|{\bf v}_i\}{\bf r}$ lie in the same cell. The action of any $\{R|{\bf v}\}\in G$ on an arbitrary ${w}^{\bf n}_{{\bf r}_i,s}$ can be decomposed in: a lattice translation, a change of the center and a local transformation. Mathematically, this is expressed as \begin{eqnarray} O\{R|{\bf v}\}\,(O\{E|{\bf n}\}\,O\{R_i|{\bf v}_i\}\, {w}^{\bf 0}_{{\bf r},s})\; =\cr O\{E|\{R|{\bf v}\}({\bf n}+{\bf r}_i)-{\bf r}_j\}\, O\{R_j|{\bf v}_j\}\, O\{\bar{R}|\bar{\bf v}\}\,{w}^{\bf 0}_{{\bf r},s} \end{eqnarray} where $\{R_j|{\bf v}_j\}\in G/G_r$ and $\{\bar{R}|\bar{\bf v}\}\in \hat{G}_{\bf r}$ are univocally determined. Together with Eq.~\ref{local-transformation}, this expression defines the band representation and, by using the inverse of Eq.~\ref{eq:w_def}, it can be written in the basis of Bloch modes. We obtain \begin{eqnarray} O\{R|{\bf v}\}\,\tilde{u}^{\bf k}_{{\bf r}_i,s}\; =\cr \exp{(-i\,R{\bf k}\,(\{R|{\bf v}\}{\bf r}_i-{\bf r}_j))}\; \sum_{h=1}^{d_{\tau}}\,D^{\tau}_{hs}(\{\bar{R}|\bar{\bf v}\})\, \tilde{u}^{R{\bf k}}_{{\bf r}_j,h} \label{band-representation} \end{eqnarray} By examining the representations this equation defines in the high symmetry k-stars, it can be easily checked whether the band representation $({\bf r},\tau)$ is equivalent to the band symmetry we want to describe.~\cite{bacry-michel-zak} Once a convenient $({\bf r},\tau)$ set is chosen, Eq.~\ref{band-representation} fixes the requirements on Bloch modes so that they lead to symmetry adapted local modes $w^{\bf n}_{{\bf r}_i,s}$. For pure translations, Eq.~\ref{band-representation} reduces to Bloch theorem. Point symmetry determines the transformation properties of the $d_{\tau}d_r$ Bloch modes in each k-point and establishes the relationship of these with those in the rest of the k-star. However, Eq.~\ref{band-representation} does not determine the form of the $M^{\bf k}$ matrices completely. For instance, in a general k-star ($\bar{G}^{\bf k}=\{E\}$) no condition is imposed on the choice of Bloch modes in a representative ${\bf k}$, though, once this is done, the modes in the rest of the star are fixed. This is the freedom we use in our construction procedure to get the localization of the modes.
\section{Introduction} \label{intro} Correlations of two nonidentical particles at small relative velocities are - due to final state interactions - sensitive to the space-time structure at freeze-out. Usually, the correlation is used to extract the size of the source and the time duration of the emission \cite{Podgoredsky73,Kopylov74,Koonin77,Pratt84,Pratt86,Pratt87,Boal90}. However, it contains also information on the emission time differences of the two particles. Gelderloos and Alexander proposed to construct velocity difference spectra at small relative angles \cite{Gelderloos94}. Comparing these spectra with results of trajectory model calculations they were able to infer the emission order and the time intervals between the emission of the two particles \cite{Gelderloos95}. If the faster particle of a pair approaches and passes the other one from behind the pair experiences a stronger final-state interaction than in the case that the faster particle has started in front of the slower one. Thus, the ratio of the correlation functions with the relative velocity parallel and anti-parallel to the pair velocity is related to the formation sequence. For heavy-ion collisions of a few hundred MeV per nucleon mainly nucleons and light composite particles with charge number $Z \le 2$ are emitted, and one expects relative short time differences. For these light ejectiles classical trajectory calculations are not accurate as quantum effects dominate at momenta larger than $\hbar/a_0$ where $a_0$ is the Bohr radius. Very recently, Lednick\'{y} et al. \cite{Lednicky96} have proposed to use the above sensitivity of the correlation function to directly measure the space-time differences in the emission of particles of different types. This is possible since the wave functions for nonidentical particles are asymmetric with respect to the forward and backward direction of the relative momentum. First theoretical \cite{Voloshin97} and experimental work \cite{Miskowiec98} concentrated on the determination of the asymmetry of particle production of proton-pion pairs. In the present work the proposed method is applied to measure the pair-wise space-time differences of different light charged particles. Applying directional cuts on all relative-velocity correlation functions of nonidentical particles allows one to determine - even with a certain redundancy - the whole sequence of space-time emission points of p, d, t, $^3$He, and $\alpha$ particles. The two-particle correlations are sensitive to the spatial separation of the two particles at the time when the second of them freezes out. This separation is the sum of the spatial separation between the two freeze-out points, and the time separation, multiplied by the velocity of the first particle. A correlation measurement does not allow to disentangle these two components. Therefore, in the following investigation we will discuss two cases, first that all particles are emitted from the same region and second that they emerge at the same time instant. \section{The experiment} \label{experiment} \subsection{Detector setup} \label{setup} The experiment has been performed at the heavy-ion synchrotron SIS at GSI Darmstadt. Targets of 1~\% interaction thickness of $^{96}$Ru and $^{96}$Zr have been irradiated by $^{96}$Ru and $^{96}$Zr ions of 400$A\cdot$MeV beam energy. In order to get sufficient statistics, the data of all target-projectile combinations have been used. The original aim of the utilization of target and projectile nuclei with equal mass but different isospin content was to answer the question whether the colliding nuclear system attains a full thermo-chemical equilibrium during the collision process. The first experimental results reveal substantial transparency effects in phase space regions already slightly apart of midrapidity \cite{Leifels98}. The present analysis uses a subsample of the data, taken with the outer Plastic Wall/Helitron combination of the FOPI detector system \cite{NIM}. The Plastic Wall delivers - via energy loss vs. time-of-flight (TOF) measurement - the nuclear charge $Z$ and the velocity ${\bf \beta}$ of the particles. The Helitron gives the curvature (which is a measure of the momentum over charge $(p/Z)$) of the particle track in the field of a large superconducting solenoid. Since the momentum resolution of the Helitron is rather moderate, this detector component serves for particle identification only. The mass $m$ is determined via $m c=(p/Z)_{Hel}/(\beta \gamma/Z)_{PlaWa}$, where $\gamma=(1-\beta^2)^{-1/2}$. The Plastic Wall and the Helitron have full overlap only for polar angles between 8.5 degrees and 26.5 degrees. The corresponding flight paths amount to 450~cm and 380~cm, respectively. Monte-Carlo simulations have been performed in order to study the influence of the finite detector granularity and of the TOF and position resolutions on the velocity and finally on the proton momentum. The resolution of both quantities is governed by the TOF resolution, which is $\sigma_{TOF} = 80$~(120)~ps for short (long) scintillator strips located at small (large) polar angles \cite{NIM}. The detector granularity delivers a negligible contribution to the velocity resolution \cite{Kotte97}. Thus, between midrapidity ($y_{cm}=0.447$, $\beta_{cm}=0.419$) and projectile rapidity ($y_{proj}=0.894$, $\beta_{proj}=0.713$) the velocity can be determined with a precision of $\sigma_{\beta}/\beta \simeq 0.4 \% - 0.8 \%$. Finally, from the velocity other kinetic quantities like the velocity ${\bf v}^{cm}$ and the particle momentum ${\bf p}^{cm}$ after transformation into the c.m. system of the colliding nuclei are deduced. \subsection{Event classification} \label{centrality} \begin{figure}[t] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.1\linewidth \epsffile{fig1.eps} } \caption{ Two-dimensional distribution of yields $d^2\sigma/dp_t dy$ of p, d, t, $\alpha$ particles in the $p^0_t-y^0$ plane for central reactions selected by a 8~\% cut on large charged particle multiplicities. Target and projectile rapidities are given by $y^0=$~-1 and +1, respectively. The full lines are levels of constant yield of 20, 40, 60, and 80~\% of the maximum value. Dashed lines represent the polar angle limits at 8.5 and 26.5 degrees. \label{phase_space} } \end{figure} About $4\cdot 10^6$ central collisions are selected by demanding large charged-particle multiplicities to be measured in the outer Plastic Wall. The corresponding integrated cross section comprises about 8~\% of the total cross section. For this centrality class one would expect - within a geometrical picture - an average impact parameter of about 2~fm. Simulations which we have performed with the IQMD model \cite{Bass} predict an average impact parameter of about 2.5~fm. For different particles with mass number $A_{clus}$ Fig.\,\ref{phase_space} shows the phase space coverage of the detector components, outer Plastic Wall/Helitron, in the transverse momentum vs. rapidity plane for events selected by this centrality condition. Here, $p_t^0=(p_t/A_{clus})/(p_{proj}/A_{proj})_{cm}= (\beta_t \gamma)/(\beta \gamma)^{proj}_{cm}$ and $y^0=(y/y_{proj})_{cm}=(y/y_{cm}-1)$ are the normalized transverse momentum and rapidity, respectively. Both observables are related to the corresponding projectile quantities in the c.m. frame of the colliding nuclei (with $(\beta \gamma)^{proj}_{cm}=0.462$ for 400 $A\cdot$MeV beam energy). It is obvious that for central collisions the Plastic Wall preferentially measures midrapidity particles with small velocities in the c.m. system ($\langle v^{cm} \rangle \simeq 0.25~\mbox{c}- 0.30~\mbox{c}$). In previous investigations of central Au+Au collisions between 100 and 400 $A\cdot$MeV beam energy it was found that the correlation function of pairs of intermediate mass fragments (IMF) is strongly affected by the collective directed sideward flow of nuclear matter \cite{Kaempfer93,Kotte95}. This directed sideflow causes an enhancement of correlations at small relative momenta. The enhancement results from mixing of differently azimuthally oriented events; it vanishes if the events are rotated into a unique reaction plane, which is determined by the standard transverse momentum analysis \cite{Odyniec}. This procedure is allowed because the geometrical acceptance and the detector efficiency are azimuthally symmetric. Thus, the technique of event rotation is applied also to the present data in order to prevent that such artificial correlations are introduced into the reference momentum distribution of the correlation function (cf. Sect.\,\ref{corr_fun}). \subsection{Correlation function}\label{corr_fun} Let $Y_{12}({\bf p}_1, {\bf p}_2)$ be the coincidence yield of pairs of particles having momenta ${\bf p}_1$ and ${\bf p}_2$. Then the two-particle correlation function is defined as \begin{equation} 1 + \mbox{R}({\bf p}_1, {\bf p}_2) = {\cal N} \, \frac{\sum _{events,pairs} Y_{12}({\bf p}_1, {\bf p}_2)} {\sum_{events,pairs} Y_{12,mix}({\bf p}_1, {\bf p}_2)}. \end{equation} The sum runs over all events fulfilling the above mentioned global selection criterion and over all pairs satisfying certain conditions given below. Event mixing, denoted by the subscript ''mix'', means to take particle 1 and particle 2 from different events. ${\cal N}$ is a normalization factor fixed by the requirement to have the same number of true and mixed pairs. The statistical errors of all the correlation functions presented below are governed by those of the coincidence yield, since the mixed yield is generated with two orders of magnitude higher statistics. The correlation function (1) is then projected onto the relative momentum ${\bf q}$, \begin{equation} {\bf q}=\mu {\bf v}_{12} =\mu ({\bf v}^{cm}_1 - {\bf v}^{cm}_2 ). \label{defq} \end{equation} Here, ${\bf v}^{cm}_{i}$ are the particle velocities calculated in the c.m. system of the colliding nuclei and $\mu=(m_1 m_2)/(m_1+ m_2)$ is the reduced mass of the pair. Besides the above described global event characteristics we use gate conditions on the angle $\zeta$ between ${\bf q}$ and the c.m. sum momentum of the particle pair ${\bf P}^{cm}_{12} ={\bf p}^{cm}_1 + {\bf p}^{cm}_2$ and on the pair velocity $V=\vert {\bf P}^{cm}_{12}\vert/(m_1+m_2)$. In order to exploit the full available statistics, two complement types of longitudinal correlation functions are generated. The forward and backward correlation functions are defined by cuts on the angle $\zeta$, $\cos{\zeta}>0$ and $\cos{\zeta}<0$, respectively. This choice selects pairs with the longitudinal velocity component $v_L$ (projection onto the pair velocity) of particle~1 being greater or smaller than the corresponding value of particle~2: \begin{equation} \mbox{R}^+(q)=1+\mbox{R}(q,\cos{\zeta>0})=1+\mbox{R}(q,v_{L,1}>v_{L,2}) \end{equation} \begin{equation} \mbox{R}^-(q)=1+\mbox{R}(q,\cos{\zeta<0})=1+\mbox{R}(q,v_{L,1}<v_{L,2}) \end{equation} From the velocity resolution as estimated in Sect.\,\ref{experiment} the corresponding $q$ resolution is deduced. It is expected to amount to $\delta q=\mu \delta v_{12}$, where $\delta v_{12}=\sqrt{2\langle(\delta v^{cm})^2\rangle}=(0.006\pm 0.002)$~c. \section{Analysis} \label{analysis} \subsection{Correlations from final-state interaction} \label{model} The correlation function of two particles 1 and 2 which move with a pair velocity ${\bf V}$ is \cite{Koonin77,Bauer1} \begin{eqnarray} \nonumber 1+R({\bf p_1},{\bf p_2})&=& \int d t_1 d t_2 d{\bf r}_1 d{\bf r}_2 \rho_1({\bf V},{\bf r}_1,t_1) \rho_2({\bf V},{\bf r}_2,t_2) \times \\ &&\vert\Psi_{\bf q}({\bf r}_1-{\bf r}_2-{\bf V}(t_1-t_2)) \vert^2, \label{defC} \end{eqnarray} where the density $\rho_{1,2}({\bf r},t)$ describes the probability to find particle 1 or 2, respectively, at time $t$ and spatial coordinate ${\bf r}$ from which they are emitted (freeze-out configuration). The wave function $ \Psi_{{\bf q}} $ describes the relative motion of the two particles with momenta ${\bf p}_1$ and ${\bf p}_2$. Since the interaction with the source is neglected this wave function depends only on the relative coordinate ${\bf r}= {\bf r}_1-{\bf r}_2$ and on the relative momentum ${\bf q}$ defined in (\ref{defq}). Here, we assume that the source functions $\rho_{1,2}$ do not depend on the velocity and have a Gaussian shape in space and time characterized by radius parameters $R_{1,2}$ and emission times $\tau_{1,2}$ while the duration of the emission is given by $\tau_0$: \begin{equation} \rho_{1,2}({\bf r},t)=\frac{1}{4 \pi^2 R_{1,2}^3 \tau_0} \exp[-\frac{r^2}{2 R_{1,2}^2} -\frac{(t-\tau_{1,2})^2}{2 \tau_0^2}]. \label{defemiss} \end{equation} The simple form of Eq.\,(\ref{defemiss}) allows us to integrate over the center-of-mass coordinate $ (m_1 {\bf r}_1 + m_2 {\bf r}_2) / (m_1+m_2)$ and the time variables leading to \begin{equation} 1+R({\bf p_1},{\bf p_2})= \int d^3r S({\bf r},{\bf V})\vert\Psi_{{\bf q}}({\bf r}) \vert^2 \label{simpleS} \end{equation} with \begin{eqnarray} \nonumber S({\bf r},{\bf V}) &=& \frac{1}{(4\pi)^{3/2}R_0^2 D} \exp\Big(-\frac{1}{4D^2}\Big[({\bf V}\Delta \tau+{\bf r})^2 \\ &&\phantom{\frac{1}{(4\pi)^{3/2}R_0^2 D}} + \frac{\tau_0^2}{R_0^2}(r^2 V^2 - ({\bf r \cdot V})^2) \Big] \Big), \label{source} \end{eqnarray} where we have introduced the time difference $\Delta \tau_{12}$ and an effective radius $D$ by \begin{equation} \Delta \tau_{12}=\tau_1-\tau_2, \quad D^2=R_0^2+(V \tau_0)^2, \quad R_0^2=\frac{1}{2}(R_1^2+R_2^2). \quad \label{dtau} \end{equation} For a finite emission time difference $\Delta \tau_{12}$ the source function $S$ depends on the sign of the scalar product ${\bf V\cdot r}$. This asymmetry with respect to ${\bf r}$ is transferred to the correlation function (\ref{simpleS}) if the wave function $\Psi_{\bf q}$ contains terms of an odd power of the scalar product ${\bf q \cdot r}$ which is possible for nonidentical particles. The wave function $ \Psi_{{\bf q}}$ is generated by partial wave expansion technique. To incorporate the spin degrees of freedom we use either spin-spin coupling (for p-p, d-p, t-p, $^3$He-p, d-d, t-d, $^3$He-d, and $^3$He-t correlations) or l-s coupling (for $\alpha$-p, $\alpha$-d, $\alpha$-t, and $\alpha$-$^3$He correlations). Thus, the partial waves are classified by angular momentum and either total spin $s=s_1+s_2$ or $j=l+s_2$. There is also an option to include partial waves with total angular momentum $J=l+s$ to describe dominant resonances. The corresponding radial Schr\"odinger equations are solved using the Coulomb potential and Woods-Saxon potentials \begin{equation} V(r)=\frac{V_{ws}}{1+\exp[-(r-R_{ws})/a_{ws}]}. \end{equation} For each partial wave, the parameters $R_{ws}$, $a_{ws}$, V$_{ws}$ are chosen such that the phase shifts are reproduced. As already mentioned in ref. \cite{Boal90} it is important to find potentials which generate the correct dependence $d \delta / d q$ since the derivative of the phase shift is the relevant quantity in determining whether correlations are suppressed or enhanced. In most of the cases we make use of the parameters already obtained in refs.\,\cite{Boal90,Boal86,Jennings86}. (Note that for $q$ values approaching a certain resonance positions $q_i$ the phase shift has to amount 90 degrees.) For partial waves $l>3$ the nuclear part of the potential is unimportant and has been neglected. In the case of p-p correlations the wave function has been generated using the Reid soft-core potential like in the standard Koonin model \cite{Koonin77}. The wave function $ \Psi_{{\bf q}}$ in Eq.\,(\ref{defC}) is the projection of the many body wave function of the two interacting clusters on their relative coordinates. The assumption made above that this wave function can be replaced by the partial waves could be violated for small distances $\vert {\bf r}_1 -{\bf r}_2 \vert$ where many body effects are most important. Consequently, the potentials which are obtained by fitting the scattering phases might not be the optimum choice to produce the correlation function. E.g., the d-p correlation functions shown in Fig.\,\ref{time_sequence_prot_deut} are too steeply increasing with relative momentum $q$. The agreement to the measurement is largely improved using the potential depth of 6~MeV instead of the value of -13~MeV given in ref.\,\cite{Boal90} for the $s=3/2$, $l=1$ partial wave. However, it is important to note that by changing this value the ratio $\mbox{R}^+/\mbox{R}^-$ remains essentially unchanged. Therefore, we have decided to use potentials which are compatible with scattering phases avoiding the ambiguity which could arise by searching for new parameters in the very large parameter space. \subsection{Effect of radial flow} \label{flow_correct_radius} From Eqs. (\ref{source}) and (\ref{dtau}) we observe that differences of the source extensions do not enter in the correlation function. This fact is a consequence of the isotropic distribution of the emission points which is independent of the particle momentum. However, in central nucleus-nucleus collisions a considerable amount of the bombarding energy is converted into flow energy. This collective expansion of nuclear matter sets in after the compression phase and can be observed as a decrease of the slope of the kinetic energy spectra. It causes a strong correlation between particle momenta and emission points. Therefore, we investigate in this section the influence of the flow on both the source radius and the emission time. For this purpose we introduce into the emission function (\ref{defemiss}) a radial flow velocity. There is good experience \cite{Reisdorf97} with the "nuclear Hubble scenario" which suggests a linear velocity profile. Such an assumption has furthermore the advantage that the function $S({\bf r},{\bf V})$ can be calculated analytically when using Gaussian density profiles. Thus, we write for particles with different mass numbers $A_{1,2}$ \begin{eqnarray} \nonumber \rho_{1,2}({\bf v},{\bf r},t)&=& N \exp[-\frac{r^2}{2 R_{1,2}^2} -\frac{(t-\tau_{1,2})^2}{2 \tau_0^2}\\ &&\phantom{N \exp[ } -\frac{A_{1,2} m_0 ({\bf v}-\eta{\bf r})^2}{2 T}], \label{newrho} \end{eqnarray} where $N$ is a normalization factor which ensures that the density is normalized to unity. In (\ref{newrho}) the velocities of the particles are thermally distributed, characterized by the temperature $T$, around a radial flow velocity defined by the scaling parameter $\eta$. The "nuclear Hubble constant" $\eta$ is connected with the radial flow energy per nucleon \begin{equation} \epsilon_{flow}=\frac{m_0 \eta^2}{2}\langle r^2\rangle= \frac{m_0 \eta^2}{2} 3 R_0^2. \label{flow_energy} \end{equation} The brackets imply averaging over the Gaussian density distribution, and the quantities $m_0$ and $R_0$ represent the nucleon rest mass and the mean radius, respectively. Repeating the integration of Eq.\,(\ref{defC}) with the new functions $\rho_{1,2}$ one arrives at a similar relation for the source function $S$ as Eq.\,(\ref{source}), however the parameters $R_0$, $D$ and $\Delta \tau_{12}$ of Eq. (\ref{dtau}) have to be replaced by quantities indicated by an asterisk: \begin{equation} R^*_0=\sqrt{ \frac{1}{2} (\frac{R_1^2}{1+\epsilon \tilde r_1^2 A_1} + \frac{R_2^2}{1+\epsilon \tilde r_2^2 A_2} )}\,, \label{radius_flow_correction} \end{equation} \begin{equation} D^* = \sqrt{(R^*_0)^2 + (V\tau_0)^2} \label{d_star} \end{equation} and \begin{equation} \Delta \tau_{12}^* = \tau^*_1 - \tau^*_2 = \Delta \tau_{12} + \Delta \tau_{12}^{flow} \label{delta_tau_star} \end{equation} with \begin{equation} \tau^*_{1,2} = \tau_{1,2} - \frac{\tilde r_{1,2}^2 A_{1,2}}{1+\epsilon \tilde r_{1,2}^2 A_{1,2}} \sqrt{\epsilon\frac{m_0}{T}R_0^2}\,, \label{tau_star} \end{equation} and \begin{equation} \Delta \tau_{12}^{flow}= -\Big(\frac{\tilde r_1^2 A_1}{1+\epsilon \tilde r_1^2 A_1} -\frac{\tilde r_2^2 A_2}{1+\epsilon \tilde r_2^2 A_2} \Big) \sqrt{\epsilon\frac{m_0}{T}R_0^2}\,. \label{delta_tau_flow} \end{equation} Here, we use the abbreviations $\tilde r_{1,2}=R_{1,2}/R_0$, and $\epsilon=\epsilon_{flow}/\epsilon_{therm}$ is the ratio of the radial flow energy $\epsilon_{flow}$ and the energy of the random thermal motion $\epsilon_{therm}=\frac{3}{2}T$. From Eq.\,(\ref{radius_flow_correction}) one finds that the apparent source radius decreases monotonously with increasing energy ratio and particle mass. These observations do well compare with recent results of the investigation of the sensitivity of the proton-proton correlation to collective expansion in central Ni+Ni collisions at $1.93\, A\cdot$GeV beam energy \cite{Kotte97}. The essential result is that even if the two particles are produced at the same source region, $R_1=R_2$, and at the same time, $\tau_1 = \tau_2$, the flow still causes an effective time difference $\Delta \tau_{12}^* = \Delta \tau_{12}^{flow}$ if the two particle masses are different. This can be understood as follows: For a given pair velocity ${\bf V}$ the heavier partner 1 with the smaller thermal velocity moves approximately with the flow velocity. This means that at freeze-out the heavier particle is located in the region around ${\bf r}_1={\bf V}/\eta$. Since the position of the lighter partner with the larger thermal velocity is Gaussian distributed around the center of the source, its mean location ${\bf r}_2$ is behind its heavier partner. Therefore, to have a strong final state interaction the lighter needs to start earlier. If they start however at the same time it seems as if the heavier particle would have started earlier, $\Delta \tau_{12}^* < 0$ in agreement with Eqs.\,(\ref{delta_tau_star}) and (\ref{tau_star}). Furthermore, temporal and spatial differences in the source distribution add up to the effective emission times $\tau^*_i$ in Eq.\,(\ref{tau_star}) quite independently. Therefore, measuring the quantity $\Delta \tau_{12}^*$ alone will not allow us to disentangle the two different contributions. In the following analyses we will therefore differ two cases, namely first that all particles are emitted from the same source size ($R_i = R_0$) and second that they are emitted at the same time ($\tau_i = 0$) instant. In our calculations which take the collective expansion into account we use an energy ratio of $\epsilon=1$ and a temperature of $T=37$~MeV. These values are compatible with values obtained in systematic flow studies \cite{Reisdorf97} of central Au+Au collisions in case of 400 $A\cdot$MeV beam energy. It should be mentioned that the flow correction $\Delta \tau^{flow}_{12}$ changes only slowly with the energy ratio $\epsilon$ because the temperature $T$ in Eq.\,(\ref{delta_tau_flow}) is correlated with the flow energy $\epsilon_{flow}$ due to the requirement of energy conservation. Thus, even if for the present Ru+Ru system the ratio $\epsilon$ would be smaller than for the Au+Au system, one would find that the correction changes only marginally. E.g. a drastic reduction of $\epsilon$ by a factor of two leads to typical changes of $\Delta \tau_{12}^{flow}$ by about 20\% only. \section{Results and comparison with model} \label{results} \subsection{The source extension}\label{source_radius} \begin{figure}[b] \centering \mbox{ \hspace*{-5bp} \epsfxsize=0.55\linewidth \epsffile{fig2a.eps} \hspace*{-18bp} \epsfxsize=0.55\linewidth \epsffile{fig2b.eps} } \caption{Left panel: The experimental correlation function of proton pairs from central trigger events (dots). The hatched area indicates the unreliable region which may be contaminated by doubly counted scattered particles. The full line represents the model prediction for an emission from a Gaussian source of zero lifetime and radius $R_0=5.7$~fm. The dashed and dotted lines are the corresponding results for source radii which differ from the optimum one by -0.3~fm and +0.3~fm, respectively. \protect\\ Right panel: The transverse (dots) and longitudinal (squares) two-proton correlation functions compared to the model results (dashed and dotted lines) for finite lifetime $\langle V \rangle \tau_0=2.2$~fm and Gaussian radius $R_0=5.4$~fm. \label{pcor_ru400_hwb} } \end{figure} Fig.\,\ref{pcor_ru400_hwb} (left panel) shows the angle-integrated two-proton correlation function. As in previous fragment-fragment correlation analyses \cite{Kaempfer93,Kotte95} an enhanced coincidence yield at very small relative angles is observed, which is due to double counting caused mainly by scattering in the scintillator strips. This disturbing yield is reduced drastically by the requirement to match the particle hits on the Plastic Wall with the corresponding tracks in the forward drift chamber Helitron. However, mismatches of tracks and scattering processes especially at the wire planes of the chamber can give rise to a small amount of double counting, too. The remaining left-over of doubly counted scattered particles is eliminated by excluding, around a given hit, positions within a rectangular segment of azimuthal and polar angle differences $\vert \phi_1 -\phi_2 \vert <4^o$ and $\vert \theta_1 -\theta_2 \vert <2^o$. In order to keep the influence of the exclusion onto the correlation function as small as possible, the same procedure is applied to the uncorrelated background. However, GEANT simulations \cite{GEANT} have shown that at very small relative velocities $v_{12} < 0.03$~c a small bias of the correlation function cannot be excluded \cite{Kotte97}. Though ratios of correlation functions as our forward/backward relation $\mbox{R}^+/\mbox{R}^-$ are expected to be rather robust against such biases, the corresponding regions in the correlation functions are marked as hatched area and are not taken into consideration when comparing the experimental data with model predictions. The experimental p-p correlation function is compared to the model for an emission from a source of zero lifetime and true Gaussian radius $R_0$ which corresponds to an r.m.s. radius of $R_{rms}\equiv \sqrt{\langle r^2 \rangle}=\sqrt{3} R_0$. The theoretical correlation function is folded with an experimental resolution function of Gaussian shape with the dispersion $\delta(q)$ as estimated in Sect.\,\ref{corr_fun}. The apparent reduction (\ref{radius_flow_correction}) of the source radius due to radial expansion effects as described in Sect.\,\ref{flow_correct_radius} (see also ref.\,\cite{Kotte97}) is estimated to $R_0^*/R_0=1/\sqrt{1+\epsilon}= 1/\sqrt{2}$. The best agreement of experimental data and model calculations is found for $R_0=(5.7 \pm 0.3)$~fm. It is obvious that the appearance of the correlation peak at $q\simeq 20$~MeV/c allows the determination of the source radius with rather high sensitivity. This $^2$He-resonance is the result of the common action of the enhancement due to the attractive nucleonic potential and the suppression due to both the mutual Coulomb repulsion and the antisymmetrization of the wave function. The ambiguity of the space and time extents of the proton emitting source can be resolved by constructing correlation functions with the relative momentum ${\bf q}$ being either perpendicular (transverse correlation function, here defined by $\vert \cos \zeta \vert <0.5$) or parallel (longitudinal correlation function, here defined by $\vert \cos \zeta \vert >0.5$) to the pair velocity ${\bf V}$. The result is given in the right panel of Fig.\,\ref{pcor_ru400_hwb}. We find a small suppression of the transverse correlations with respect to the longitudinal ones which is consistent with model predictions for the emission from a source of finite lifetime \cite{Koonin77,Pratt87}. The best simultaneous fit around the correlation peaks ($\mbox{14 MeV/c}<q<\mbox{42 MeV/c}$) of both the transverse and longitudinal correlation functions delivers Gaussian parameters of the radius (after flow correction) and the emission duration of $R_0=5.4$~fm and $\langle V \rangle \tau_0=2.2$~fm, respectively. Obviously, the lifetime effect appears unimportant. For the determination of the emission time differences in the subsequent section the source radius is kept fixed to a value of $R_1=R_2=R_0=5.7$~fm. For simplification the duration of the emission $\tau_0$ is set to zero. A finite value leads only to a rescaling of the effective radius in Eq.\,(\ref{d_star}). \subsection{Emission time differences}\label{time_differences} Out of the ten different combinations of pairs of nonidentical particles which can be constructed from p, d, t, $^3$He, and $\alpha$ particles there exist only two correlation functions which do not contain contributions of resonance decay products. These resonance-free correlation functions of deuteron-proton and $^3$He-triton pairs are presented in Sect.\,\ref{resonance_free_correlations}. All other correlation functions contain resonance contributions of particle-unbound ground states or excited states of heavier clusters decaying into pairs of light charged particles. In case of the appearance of strong and narrow resonances in the $q$ region of interest, the corresponding theoretical correlation functions are corrected for the experimental $q$ resolution similarly as for p-p correlations. \subsubsection{Final-state interaction without resonances} \label{resonance_free_correlations} \begin{figure}[t] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.\linewidth \epsffile{fig3.eps} } \caption{ Upper panel: Forward (full dots) and backward (open dots) longitudinal experimental correlation functions of d-p pairs. The hatched area indicates the unreliable region which may be contaminated by doubly counted scattered particles. The full and dashed lines give the corresponding model predictions with the time delay of table\,\protect\ref{pair_wise_time_delays}. Lower panel: Ratio of forward/backward experimental correlation functions (open squares). The full line represents the ratio of the simulated correlation functions. The dashed and dotted lines give the model predictions for times differences deviating by -0.5~fm/c and +0.5~fm/c from the optimum one, respectively. \label{time_sequence_prot_deut} } \end{figure} \begin{figure}[b] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.\linewidth \epsffile{fig4.eps} } \caption{ The same as Fig.\,\protect\ref{time_sequence_prot_deut}, but for $^3$He-t correlations. \label{time_sequence_3he_trit} } \end{figure} Fig.\,\ref{time_sequence_prot_deut} gives the results of d-p correlations. The upper panel shows the forward (full dots) and the backward (open dots) correlation functions whereas the lower panel represents the ratio of both observables. The resonance-free correlation functions show a suppression at low relative momenta due to final-state Coulomb and nuclear interactions. Obviously, the suppression gets stronger in the case when deuterons are faster than protons (full dots in Fig.\,\ref{time_sequence_prot_deut}). This indicates that, on average, deuterons are being emitted later than protons. Quantitatively, the comparison with the model predictions yields an optimum time delay of $\Delta \tau_{\mbox{d,p}}= \tau_{\mbox{d}} - \tau_{\mbox{p}}=6.5$~fm/c (full line in the lower panel). (The dashed and dotted lines in the lower panel of Figs.\,\ref{time_sequence_prot_deut} and \ref{time_sequence_3he_trit} should give an impression how the theoretical ratio $\mbox{R}^+/\mbox{R}^-$ alters if the time delay is changed by -0.5~fm/c and +0.5~fm/c, respectively.) However, if one does not take into account the time shift due to the radial expansion of the participant zone the emission time difference would be determined as an apparent value which is much smaller (but still positive). Both the true and apparent time delays are summarized in table\,\ref{pair_wise_time_delays}. The given errors represent the typical band widths of time delay parameters around their optimum values which reproduce the experimental correlation function ratios $\mbox{R}^+/\mbox{R}^-$ reasonably well. \begin{table}[b] \vspace*{10bp} \caption{The apparent and the true values of the emission time difference as determined for all combinations of nonidentical charged-particle correlation functions. The source radii are fixed to $R_i=R_0$. \label{pair_wise_time_delays} } {\setlength{\tabcolsep}{5.5mm} \begin{tabular}{rrr} \hline light-charged & apparent & true \protect\\ particle & time delay & time delay \protect\\ combination & $\Delta \tau_{12}^*$ & $\Delta \tau_{12} = \tau_1 -\tau_2$ \protect\\ 1 - 2 & (fm/c) & (fm/c) \protect\\ \hline d - p & $ 1.7 \pm 1$ & $6.5 \pm 1$ \protect\\ t - p & $ -4.2 \pm 3$ & $3 \pm 3$ \protect\\ $^3$He - p & $4.5 \pm 1$ & $11.7 \pm 1$ \protect\\ $\alpha$ - p & $-2.6 \pm 3$ & $ 6 \pm 3$ \protect\\ t - d & $ -1.0 \pm 1$ & $1.4 \pm 1$ \protect\\ $^3$He - d & $3.0 \pm 1$ & $5.4 \pm 1$ \protect\\ $\alpha$ - d & $-2.0 \pm 1$ & $1.8 \pm 1$ \protect\\ $^3$He - t & $1.7 \pm 1$ & $1.7 \pm 1$ \protect\\ $\alpha$ - t & $ -2.9 \pm 1$ & $-1.5 \pm 1$ \protect\\ $^3$He - $\alpha$ & $4.4 \pm 1$ & $3.0 \pm 1$ \protect\\ \hline \end{tabular} } \end{table} Fig.\,\ref{time_sequence_3he_trit} shows the $^3$He-t correlation functions and the corresponding forward/backward ratio. Since the masses of the two species are practically identical, this correlation function is the only one where - in case of equal source radii - the flow correction (\ref{delta_tau_flow}) vanishes and, consequently, the true and the apparent time delays are identical. Obviously, the $^3$He particles are emitted slightly later than the tritons (about 1-3~fm/c, cf. table\,\ref{pair_wise_time_delays}). \subsubsection{Final-state interaction with resonances} \label{resonance_containing_correlations} \begin{figure}[t] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.\linewidth \epsffile{fig5.eps} } \caption{ Upper panel: Forward (full dots) and backward (open dots) longitudinal experimental correlation functions of t-p pairs. Positions of relevant resonances are marked as arrows. The hatched area indicates the unreliable region which may be contaminated by doubly counted scattered particles. The full and dashed lines give the corresponding model predictions with the time delay of table\,\protect\ref{pair_wise_time_delays}. Lower panel: Ratio of forward/backward experimental correlation functions (open squares). The full line represents the ratio of the simulated correlations. \label{time_sequence_prot_trit} } \end{figure} \begin{figure}[t] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.\linewidth \epsffile{fig6.eps} } \caption{ The same as Fig.\,\protect\ref{time_sequence_prot_trit}, but for $^3$He-p correlations. \label{time_sequence_3he_prot} } \end{figure} \begin{figure}[t] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.\linewidth \epsffile{fig7.eps} } \caption{ The same as Fig.\,\protect\ref{time_sequence_prot_trit}, but for $\alpha$-p correlations. The dark hatched area indicates the unreliable region which may be contaminated by doubly counted scattered particles. The light hatched area gives the region which is supposed to be strongly contaminated by secondary decays of boron isotopes. \label{time_sequence_4he_prot} } \end{figure} \begin{figure}[b] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.\linewidth \epsffile{fig8.eps} } \caption{ The same as Fig.\,\protect\ref{time_sequence_prot_trit}, but for t-d correlations. \label{time_sequence_trit_deut} } \end{figure} \begin{figure}[t] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.\linewidth \epsffile{fig9.eps} } \caption{ The same as Fig.\,\protect\ref{time_sequence_prot_trit}, but for $^3$He-d correlations. \label{time_sequence_3he_deut} } \end{figure} If a resonance contribution is dominating in the two-particle yield the ratio $\mbox{R}^+/\mbox{R}^-$ necessarily is forced to unity for relative momenta $q$ approaching the resonance value $q_i$. This is due to the fact that for pure two-body decays both particles are emitted at the same time (and position). Indeed, the experimental data show the expected behaviour (see arrow positions in Figs.\,\ref{time_sequence_prot_trit}-\ref{time_sequence_3he_4he}). Fig.\,\ref{time_sequence_prot_trit} shows the forward/backward longitudinal correlation functions of t-p pairs together with the corresponding ratio. The correlation function exhibits a broad peak which contains the contribution of the 1st excited state of $^4$He ($E^*=20.21$~MeV, $J^{\pi}=0^+$, $\Gamma=0.5$~MeV, $\Gamma_p/\Gamma=1$, $q_1=23.6$~MeV/c) as well as the 2nd ($E^*=21.01$~MeV, $J^{\pi}=0^-$, $\Gamma=0.84$~MeV, $\Gamma_p/\Gamma=0.76$, $q_2=41.0$~MeV/c) and the 3rd one ($E^*=21.84$~MeV, $J^{\pi}=2^-$, $\Gamma=2.01$~MeV, $\Gamma_p/\Gamma=0.63$, $q_3=53.4$~MeV/c). The positions of the relevant resonances are marked by arrows. In the present case only the lowest three excited states of $^4$He with widths $\Gamma<2.1 $~MeV are indicated. Other states in the energy region $E^*=23-26$~MeV ($q=70-90$~MeV/c) are much broader ($\Gamma>5$~MeV, $\Gamma_p/\Gamma\simeq 0.5$). Obviously, from the lower part of the figure one would conclude a negative time delay $\Delta \tau^*_{\mbox{t,p}}$. However, the radial flow correction (\ref{delta_tau_flow}) overcompensates this apparent time difference and leads to a positive value (cf. table\,\ref{pair_wise_time_delays}). Due to the possible contribution of higher lying resonances of $^4$He which are not taken into account in the model description, a rather large error is appended to the time delay. Fig.\,\ref{time_sequence_3he_prot} represents the results for $^3$He-p correlations. One broad maximum shows up which is related to the particle-unbound ground state of $^4$Li ($J^{\pi}=2^-$, $\Gamma=6$~MeV, $q_0=66.0$~MeV/c). An unquestionably positive time delay $\Delta \tau_{^3\mbox{He,p}}$ is deduced from the correlation function ratio $\mbox{R}^+/\mbox{R}^-$. Here, about 60~\% of the true time difference given in table\,\ref{pair_wise_time_delays} arise from the radial flow correction (\ref{delta_tau_flow}). Fig.\,\ref{time_sequence_4he_prot} gives the $\alpha$-p correlations. The origin of the broad bump is not fully understood. The main contribution results from the decay of the particle-unbound ground state of $^5$Li ($J^{\pi}=\frac{3}{2}^-$, $\Gamma=1.5$~MeV, $\Gamma_p/\Gamma=1$, $q_0=54.3$~MeV/c). Most probably, the bump contains additional contributions \cite{Pochodzalla85,ZhiYongHe97,Charity95} which cannot be separated experimentally from the $^5$Li resonance. One contribution is expected at $q\simeq 16 $~MeV/c. It corresponds to the three-body decay of $^9\mbox{B(g.s.)}\rightarrow$ p + $^8$Be(g.s.) $\rightarrow$ p + $\alpha$ + $\alpha$, where only one of the $\alpha$ particles is detected together with the proton. In addition, the four-body decay of $^{10}$B$^* \rightarrow$ p + $^9$Be(1.69 MeV) $\rightarrow$ p + n + $^8$Be(g.s.) $\rightarrow$ p + n + $\alpha$ + $\alpha$ can contribute to the broad maximum as was deduced at a comparably low beam energy of 40 $A\cdot$MeV \cite{Charity95}. In the present case such a contribution cannot be ruled out since a considerable production of 0.3 boron clusters per event was found in central Au+Au reactions at 400 $A\cdot$MeV beam energy \cite{Reisdorf97}. Due to the large lifetime of the $^8$Be ground state ($\Gamma =6.8 $~eV) the protons - a priori - are emitted earlier than the $\alpha$ particles. Thus, for relative momenta $q\simeq 0 - 60 $~MeV/c one necessarily expects stronger final-state interaction if $v_L(\alpha)>v_L(\mbox{p})$ and consequently $\mbox{R}^+/\mbox{R}^-<0$. Indeed, the experimental data follow the predicted trend. However, since the model description does not incorporate two-stage decays, a reliable time difference can only be extracted for $q$ values well above the resonance $q_0$. Keeping in mind that the sensitivity of the correlation function to a variation of the model parameter $\Delta \tau_{12}$ decreases for increasing $q$, the deduced value is affected with a rather large error. Thus, the time difference derived from the $\alpha$-p correlation function will enter with negligible weight into the procedure used to determine the emission-time sequence of the different particle species (cf. Sect.\,\ref{discussion}). Two resonances affect the t-d correlation function (Fig.\,\ref{time_sequence_trit_deut}). The first one which is not resolved experimentally corresponds to the 16.75~MeV excited state of $^5$He ($J^{\pi}=\frac{3}{2}^+$, $\Gamma=76$~keV, $q_1=10.8$~MeV/c). The second one is due to the state at 19.8~MeV ($J^{\pi}=(\frac{3}{2},\frac{5}{2})^+$, $\Gamma=2.5$~MeV, $q_2=83.5$~MeV/c). From the lower part of the figure it is obvious that the apparent time delay is close to zero. Indeed, the model fits well the data for a true emission time difference (cf. table\,\ref{pair_wise_time_delays}) which - to a large extent - is due to the contribution of the radial flow correction (\ref{delta_tau_flow}). The $^3$He-d correlations in Fig.\,\ref{time_sequence_3he_deut} appear very similar to the t-d correlations. The correlation function contains one strong peak due to the 16.66~MeV state of $^5$Li ($J^{\pi}=\frac{3}{2}^+$, $\Gamma=0.3$~MeV, $q_1=24.8$~MeV/c). Other broader states are at 18.0~MeV ($J^{\pi}=\frac{1}{2}^+$, $\Gamma\simeq 5$~MeV, $q_2=60.3$~MeV/c) and at 20.0~MeV ($J^{\pi}=(\frac{3}{2},\frac{5}{2})^+$, $\Gamma= 5$~MeV, $q_3=90.2$~MeV/c). Only the narrow 16.66~MeV state is taken into account in the model. In contrast to the t-d correlations, we find a clearly positive apparent time delay $\Delta \tau_{^3\mbox{He,d}}^* \simeq 3$~fm/c which increases nearly by a factor of 2 due to the radial flow correction. The $\alpha$-d correlation function shown in the upper panel of Fig.\,\ref{time_sequence_4he_deut} is governed by a strong resonance due to the narrow 2.186~MeV state of $^6$Li ($J^{\pi}=3^+$, $\Gamma=24$~keV, $q_1=42.2$~MeV/c). Another state is at 4.31~MeV ($J^{\pi}=2^+$, $\Gamma=1.7$~MeV, $q_2=84.2$~MeV/c). Obviously, from the lower part of the figure one would conclude a negative time delay $\Delta \tau^*_{\alpha,\mbox{d}}$. However, for the present case the radial flow correction (\ref{delta_tau_flow}) overcompensates the apparent time difference leading to a positive value (cf. table\,\ref{pair_wise_time_delays}). As a by-product, the strong $3^+$ resonance of $^6$Li in the correlation function can serve for an independent determination of the $q$ resolution. A Gaussian fit to the difference spectrum of true and normalized mixed yields in the region 20~MeV/c~$<q<$60~MeV/c delivers a dispersion of $\delta v_{12} =\delta q / \mu = 7.1~\mbox{(MeV/c)}/\mu=0.0057$~c in good agreement with the estimate given in Sect.\,\ref{corr_fun}. \begin{figure}[t] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.\linewidth \epsffile{fig10.eps} } \caption{ The same as Fig.\,\protect\ref{time_sequence_prot_trit}, but for $\alpha$-d correlations. \label{time_sequence_4he_deut} } \end{figure} \begin{figure}[b] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.\linewidth \epsffile{fig11.eps} } \caption{ The same as Fig.\,\protect\ref{time_sequence_prot_trit}, but for $\alpha$-t correlations. \label{time_sequence_4he_trit} } \end{figure} \begin{figure}[t] \vspace*{-17bp} \centering \mbox{ \epsfxsize=1.\linewidth \epsffile{fig12.eps} } \caption{ The same as Fig.\,\protect\ref{time_sequence_prot_trit}, but for $^3$He-$\alpha$ correlations. \label{time_sequence_3he_4he} } \end{figure} The $\alpha$-t correlation function given in Fig.\,\ref{time_sequence_4he_trit} is characterized by a broad maximum due to the 4.63~MeV state of $^7$Li ($J^{\pi}=\frac{7}{2}^-$, $\Gamma=93$~keV, $q_1=83.4$~MeV/c). Other states are at 6.68~MeV ($J^{\pi}=\frac{5}{2}^-$, $\Gamma=880$~keV, $q_2=116.4$~MeV/c) and at 7.46~MeV ($J^{\pi}=\frac{5}{2}^-$, $\Gamma=89$~keV, $q_3=126.7$~MeV/c). Only the 4.63~MeV state is taken into account in the model. The deduced time delay $\Delta \tau_{\alpha,\mbox{t}}$ is apparently negative; it is reduced by about 50~\% when taking into account the radial flow correction (cf. table\,\ref{pair_wise_time_delays}). The $^3$He-$\alpha$ correlation function in Fig.\,\ref{time_sequence_3he_4he} is dominated by the 4.57~MeV state of $^7$Be ($J^{\pi}=\frac{7}{2}^-$, $\Gamma=175$~keV, $q_1=98.0$~MeV/c). Here, both the deduced apparent and true time delays $\Delta \tau_{^3\mbox{He},\alpha}$ are found positive (table\,\ref{pair_wise_time_delays}) and differ only by about 1.5~fm/c. Similarly to the $\alpha$-t correlations the small flow correction is a result of the small relative mass difference of the particles. \subsection{Discussion}\label{discussion} Finally, the redundancy of the ten different pair-wise time differences obtained above allows one to fix rather reliably the emission-time sequence of the light charged particles. The optimum sequence is derived from a least-squares solution of the set of linear equations for the time delays given in table\,\ref{pair_wise_time_delays}. (Note that, the deduced $\chi^2$ per degree of freedom of 1.1 does not carry quantitative information on the quality of the regression since the errors of the individual time delays are not real standard deviations.) The different contributions are weighted by the inverse squares of the given errors. If one excludes from the fit a few of the time differences (e.g. those which are affected with large errors like in the case of $\alpha$-p and t-p correlations) the result changes only marginally. For a fixed source radius, the most probable time order of the emission is found as follows (cf. table\,\ref{time_sequence_p_d_a_t_3he}): On the average, protons are emitted first whereas $^3$He particles are emitted last (after about 11~fm/c). The other particles show up in between. Deuterons are emitted about 6~fm/c after protons. After that, the $\alpha$ particles are emitted and then the tritons follow. However, the differences between the emission times of these species are found in the order of 1~fm/c only (note: 3~fm/c~=~10$^{-23}$~s). \begin{table}[b] \vspace*{10bp} \caption{The 2nd row gives the emission times (relative to that of the protons) for fixed source radii $R_i=R_0$ of d, t, $^3$He, and $\alpha$ particles as derived from a weighted regression of the ten linear equations for the pair-wise time delays given in table \,\protect\ref{pair_wise_time_delays}. The 3rd row gives the complementary information of different source radii (normalized to that of the proton source $R_{\mbox p}=R_0$) if the emission takes place at the same time instant. \label{time_sequence_p_d_a_t_3he} } {\setlength{\tabcolsep}{1.4mm} \begin{tabular}{ ccccc } \hline \protect\\ \vspace*{6bp} & d & $\alpha$ & t & $^3$He \protect\\[1bp] \hline \protect\\ \vspace*{6bp} {\large $\frac{ \tau_i- \tau_{\small \mbox p}}{fm/c}$ } & $6.3\pm 0.8$ & $7.7\pm 0.9$ & $8.5\pm 0.9$ & $11.1\pm0.8$ \protect\\[1bp] \hline \protect\\ \vspace*{6bp} $R_i/R_{\mbox p}$ & $0.63\pm 0.04$ & $0.54\pm 0.05$ & $0.53\pm 0.05$ & $0.44\pm 0.04$ \protect\\[1bp] \hline \protect\\ \end{tabular} } \end{table} As a consequence of the duality of space and time coordinates, the time delays can be transformed into position differences. When assuming a common emission time for all particles, the time sequence translates into the emission from different source radii $R_i$ according to Eq.\,(\ref{tau_star}) as a consequence of the radial flow. A very similar procedure as described above for the emission times at unique source size leads to the average radii for the emission at the same time instant. The results are summarized in the 3rd row of table\,\ref{time_sequence_p_d_a_t_3he}. Now, the protons, which are emitted earliest in the time-ordered picture, would come from the most extended source ($R_{rms}=\sqrt{3} R_0 \simeq 10$~fm) whereas the clusters (with the exception of $\alpha$ particles, which obviously play a special role) are emitted from source radii which decrease with increasing mass. Thus, the size of the emission sources of the d, $\alpha$, t and $^3$He particles would be about 63\%, 54\%, 53\% and 44\% of that of the protons, respectively. This finding is in qualitative agreement with the results of an earlier investigation \cite{Petrovici} of the cluster formation process in central Au+Au collisions at 250 $A\cdot$MeV. There, the authors compare the experimental data with predictions of a model which describes the hydrodynamic isotropic expansion of an ideal nucleonic gas and the clustering by statistical disassembly. The model predicts a breakup which, with elapsing time, starts at the exterior and evolves to smaller radii. The heavier fragments are found to arise from smaller source radii than the light particles. The $\alpha$ particles do not follow completely the systematic trend established above. One reason for this violation of the emission order with increasing mass might lie in the fact that this (strongly bound) particle species, at least partially, is made of nucleons which were initially correlated either in the target or projectile nucleus \cite{Gossiaux}. Such preformed clusters would carry to some extent a memory of the entrance channel. We have studied, whether the emission time difference of protons and composite particles can be explained by calculations we have performed with the IQMD transport model \cite{Bass}. For a central ($b<4$~fm) reaction of Ru+Ru we have investigated the distribution of the time instants at which the particles pass through the surface of a sphere in coordinate space. Only particles coming from the participant zone have been selected by a cut on the c.m. polar angle $\vert \tan{\Theta_{cm}} \vert>1$. Since correlation functions are sensitive to particle pairs with small relative momenta only, in addition, we have to demand that the velocities of both particles be the same. The time distributions of all the light charged particles are found almost symmetric and exhibit - within the statistical errors - identical mean values. Alternatively, the cross check of the radial distributions of particles with equal velocities at a unique time reveals no differences of the mean radii. This finding is not surprising since most of the transport codes generate composite particles by coalescence. Indeed, such a method provides the composite particles with the same space-time distribution as the nucleons of equal velocity. \section{Summary} \label{summary} In conclusion, we have presented experimental correlation functions of nonidentical light charged particles produced in central collisions of Ru(Zr)+Ru(Zr) at 400 $A\cdot$MeV. For the first time an emission order of p, d, $\alpha$, t, and $^3$He particles has been set up by comparing correlations of particles with relative momenta parallel and anti-parallel to the center-of-mass velocity of the pair. Collective radial expansion of the participant zone leads to an apparent reduction of the source radius and to shifts of the emission times. Correcting for both effects typical time delays of a few fm/c were obtained. The deduced space-time differences of the light-charged-particle emission sources allow two complementary interpretations. If the source radius is fixed the composite particles are emitted at later times than protons. Alternatively, if the emission time is fixed, the clusters are emitted from smaller sources than protons. As a result of the duality of space and time coordinates, these two scenarios cannot be distinguished from each other.
\section{Introduction} The use of gravitational lensing statistics as a cosmological tool was first considered in detail by \citet{ETurnerOG84a}; the influence of the cosmological constant was investigated thoroughly by \citet{MFukugitaFKT92a}, building on the work of \citet{ETurner90a} and \citet{MFukugitaFK90a}. More recently, \citet[hereafter K96, and references therein]{CKochanek96a} and \citet{EFalcoKM98a} have laid the groundwork for using gravitational lensing statistics for the detailed analysis of extragalactic surveys. However, these analyses either have concentrated on a small subset of the possible cosmological models as described by the density parameter $\Omega_{0}$ and the cosmological constant $\lambda_{0}$, have used a simpler (singular) lens model or both. This analysis is the first time $\lambda_{0}$ and $\Omega_{0}$ have been used as independent parameters in conjunction with a non-singular lens model in an analysis of this type, complementing similar analyses with other emphases. (See \citet{YChengMKrauss99a} for a discussion of the importance of including a core radius.) Also, we include enough of the $\lambda_{0}$-$\Omega_{0}$ plane to avoid neglecting any possibly viable models; this also makes the comparison with a variety of other cosmological tests easier. This is especially important in light of the fact that many analyses \citep[e.g.][]{SPerlmutteretal98a,ARiessetal98a,BSchmidtetal98a, RCarlbergetal98a,CLineweaver98a,EGuerraDW98a,RDalyGW98a} are now suggesting that our universe may contain a significant cosmological constant \emph{and} be non-flat. The plan of this paper is as follows. Sect.~\ref{theory} reviews the groundwork and serves to define our notation. In Sect.~\ref{obsprior} we specify the observational data and selection functions we use and formulate prior information about the parameters $\lambda_0$ and $\Omega_0$. Sect.~\ref{calculations} describes the parametric submodels we use and the numerical computations we perform. In Sect.~\ref{results} we discuss our results and compare them with others. Sect.~\ref{conclusions} presents our summary and conclusions. \section{Probability of multiply imaged sources} \label{theory} In this section we briefly review the statistical concepts introduced in K96; this also serves to define our notation. Note that with regard to cosmogical notation we follow that of \citet{RKayserHS97a}, repeating here only 2 equations needed for discussion in this paper: the comoving spherical volume element at redshift $z$ reads \begin{equation} \mathrm{d}V = 4\pi r^2\frac{c}{H_0} \frac{\mathrm{d}z}{\sqrt{Q(z)}}, \label{eq:vol} \end{equation} where \begin{equation} \label{eq:Q} Q(z) = \Omega_0(1 + z)^3 - (\Omega_0 + \lambda_0 - 1)(1 + z)^2 + \lambda_0. \end{equation} Following the K96 approach, we assume that the light deflection properties of the gravitational lenses can be modelled with a particular type of circularly symmetric lens models with a monotonically declining radial mass profile. Such lens models generally create three images and have two critical radii on which the magnification diverges \citep[e.g.][]{PSchneiderEF92a}. It is possible to estimate the probability $p(m,z_{\mathrm{s}})$ of the event \begin{quote} \em A source at redshift $z_{\mathrm{s}}$ is triply imaged. The total apparent magnitude of the three images is $m$. The image configuration meets the selection criteria $S$ and, particularly, shows the properties $C$. \end{quote} If the outer and inner critical angular radii of the lens potential are respectively $r_+$ and $r_-$, the image magnification at radial angular position $r$ is $\mu(r)$, the total magnification of the three images of a source at angular position $y$ is $M(y)$, the functions $S(y)$ and $C(y)$ are $0|1$ valued selection functions, the comoving density of lenses of luminosity $L$ is $\mathrm{d}n/\mathrm{d}L$ and the number-magnitude counts of sources are $\mathrm{d}N/\mathrm{d}m$, then \begin{eqnarray} p(m,z_{\mathrm{s}}) & = & \frac{1}{2} \int\limits_0^{z_{\mathrm{s}}} \frac{\mathrm{d}V}{\mathrm{d}z} \int\limits_0^\infty \frac{\mathrm{d}n}{\mathrm{d}L} \int\limits_{r_-}^{r_+} r|\mu(r)|^{-1} \qquad\times\nonumber\\ & & \times\qquad \underbrace{B(m,z_{\mathrm{s}},y)S(y)C(y)} \,\mathrm{d}r \,\mathrm{d}L \,\mathrm{d}z, \label{eq:p} \end{eqnarray} where \begin{equation} \label{eq:bias} B(m,z,y) = \frac{\mathrm{d}N}{\mathrm{d}m}\{m +2.5 \log[M(y)], z\} \left[ \frac{\mathrm{d}N}{\mathrm{d}m}(m, z) \right]^{-1}. \end{equation} The critical radii, the image magnifications and the source position are functions of the lens model, the luminosity of the lens galaxy and the redshifts of the source and the lens galaxy. If the underbraced functions are dropped, Eq.~(\ref{eq:p}) yields the optical depth -- the fraction of the sky included within the caustics of all lenses between us and the sources at redshift $z_{\mathrm{s}}$. The inclusion of these functions accounts for magnification bias, survey selection effects (including what is defined as a lensing event) and allows the observed image separation to be taken into account. Equation~(\ref{eq:p}) parametrically depends on $\lambda_0$ and $\Omega_0$ through Eq.~(\ref{eq:vol}) and through the angular size distances,\footnote{In general, the angular size distances depend not only on $\lambda_{0}$ and $\Omega_{0}$ but on the degree of homogeneity in the universe as well \citep[see, e.g.,][]{RKayserHS97a}. However, in contrast to some other cosmological tests, this effect is relatively unimportant for the type of analysis performed here \citep[see, e.g.,][]{MFukugitaFKT92a}.} which are needed for calculating observable quantities from the lens model (these also depend on the source and lens redshifts). Equation~(\ref{eq:p}) additionally depends on parametric submodels required to model the lens population and the number-magnitude counts of sources. Since throughout this paper we are principally interested in $\lambda_0$ and $\Omega_0$, hereafter we refer to the submodel parameters as nuisance parameters (although technically they are on the same footing with $\lambda_{0}$ and $\Omega_{0}$, there are not of as much interest here and thus a nuisance). In principle, one could also incorporate other observables into the parametric model; the reasons for not doing so are practical. Assuming the survey selection function $S$ is known, we can numerically compute Eq.~(\ref{eq:p}) and reasonably estimate the probability $1-p(m_i,z_i)$ that the quasar $i$ is singly imaged or the probability $p(m_i,z_i,\theta_i)$ that the quasar $i$ is multiply imaged and its images (within some tolerance) are separated by $\theta_i$. If the survey data $D$ contains $M$ singly and $N$ multiply imaged quasars, we can estimate the probability of the event \begin{quote} \em In a model universe fixed by the cosmological parameters $\lambda_0$, $\Omega_0$ and the nuisance parameters $\vec{\xi}$, a multiply imaged quasar survey collects the observational data~$D$. \end{quote} by applying the parametric model (or likelihood function) \begin{eqnarray} \label{eq:lf} \ln[p(D|\Omega_0,\lambda_0,\vec{\xi})] & = & -\sum_{i=1}^{M} p(m_i,z_i) \nonumber\\ & & + \sum_{j=1}^{N} \ln[p(m_j,z_j,\theta_j)], \end{eqnarray} where the logarithm $\ln[1-p(m_i,z_i)]$ was expanded to first order. We can combine surveys of different objects by adding the logarithms of the likelihood functions for the individual surveys, and can combine surveys containing the same objects by applying their joint selection function. In Bayesian theory the model parameters $\lambda_0$, $\Omega_0$, $\vec{\xi}$ are regarded as random quantities with known joint prior probability density function $p(\lambda_0,\Omega_0,\vec{\xi})$. Applying Bayes's theorem, the appropriate posterior probability distribution given the observational data $D$ is \begin{equation} p(\lambda_0,\Omega_0\vec{\xi}|D) = p(D|\lambda_0,\Omega_0,\vec{\xi}) \otimes p(\lambda_0,\Omega_0,\vec{\xi}), \end{equation} where the operation `$\otimes$' denotes multiplication followed by normalisation. Marginalising the nuisance parameters \begin{equation} \label{eq:marg} p(\lambda_0,\Omega_0|D) = \int\limits p(\lambda_0,\Omega_0,\vec{\xi}|D) \,\mathrm{d}\vec{\xi}. \end{equation} yields the (marginal) posterior probability density function for the parameters $\lambda_0$ and $\Omega_0$. In the limit where all nuisance parameters take a precise value, $\vec{\xi}=\vec{\xi}_0$, the joint prior probability density function $p(\lambda_0,\Omega_0,\vec{\xi})$ factorises into $p(\lambda_0,\Omega_0)$ and a delta distribution $\delta(\vec{\xi}-\vec{\xi}_0)$, and Eq.~(\ref{eq:marg}) simplifies to \begin{equation} \label{eq:simplemarg} p(\lambda_0,\Omega_0|D) = p(D|\lambda_0,\Omega_0,\vec{\xi}_0) \otimes p(\lambda_0,\Omega_0). \end{equation} On the basis of Eq.~(\ref{eq:marg}) or Eq.~(\ref{eq:simplemarg}), we can calculate confidence regions for two parameters or perform further marginalisations and calculate mean values, standard deviations and marginal confidence intervals for one parameter. \section{Observational data and prior information} \label{obsprior} We use the observational data of the optical multiply imaged quasar surveys by \citet{DCramptonMF92b}, \citet{AJaunsenJPS95a}, \citet{CKochanekFS95a}, \citet{HYeeFT93a} and the observational data of the HST Snapshot Survey compiled by \citet{DMaozBSBDDGKMY93a}, including \mbox{\object{Q 0142-100}}, \mbox{\object{Q 1115+080}} and \mbox{\object{Q~1413+117}}. If applicable, we replace the apparent quasar V magnitude catalog data found in \citet{DCramptonMF92b}, \citet{AJaunsenJPS95a} and \citet{HYeeFT93a} with more current data from \citet{MVeronCettyPVeron96a}. We estimate the \citet{CKochanekFS95a} apparent quasar V magnitude data by adding the survey average V--R and V--I colours to the observational R and I magnitude data. Following K96, we only include quasars with redshift~$z_{\mathrm{s}}>1$. In all, our sample contains~807 singly and~5 multiply imaged quasars. The observational data of the multiply imaged quasars are summarised in Table~\ref{ta:data}. Our complete input data can be obtained from \begin{quote} \verb|http://multivac.jb.man.ac.uk:8000/ceres|\\ \verb|/data_from_papers/lower_limit.html| \end{quote} \begin{table} \caption[]{ Observational data of multiply imaged quasars contained in the sample. The magnitudes are V magnitudes unless otherwise specified. The image separations are taken from \citet{CKochanekFILMR97a}} \label{ta:data} \begin{tabular*}{\linewidth}{@{\extracolsep{\fill}}llll} \hline \hline Identifier & $m$ [mag] & $z_{\mathrm{s}}$ & $\theta$ [$\arcsec$] \\ \hline \object{Q 0142$-$100} & $17.0$ & $2.72$ & $2.2$ \\ \object{Q 1009$-$0252} & $18.1$ B & $2.74$ & $1.5$ \\ \object{Q 1115$+$080} & $16.2$ & $1.72$ & $2.2$ \\ \object{Q 1208$+$1011} & $17.9$ & $3.80$ & $0.48$ \\ \object{Q 1413$+$117} & $17.0$ & $2.55$ & $1.2$ \\ \hline \end{tabular*} \end{table} This follows K96 for purposes of comparison. Since much larger surveys (i.e.~CLASS) will be considered in a future paper, there is little point in increasing the number of lenses for its own sake. Since radio observations are considered in more detail in a companion paper \citep{PHelbigMQWBK99a}, we restrict ourselves to optical surveys in this paper. We use the \citet{DCramptonMF92b}, HST Snapshot Survey and \citet{HYeeFT93a} survey selection functions proposed in \citet{CKochanek93c}, the \citet{AJaunsenJPS95a} survey selection function at $1\farcs0$ seeing and the preliminary \citet{CKochanekFS95a} survey selection function. Before considering prior information in more detail, one must first decide which region of the $\lambda_{0}$-$\Omega_{0}$ plane is to be investigated. Clearly, this region should be defined by either exact constraints or conservative estimates, as opposed to current `best fit' values (and their errors), in order to avoid excluding any possibly viable cosmological models. Also, it is desirable for the region to be on the large side, so that in addition the sensitivity of the test (i.e.~what regions of the $\lambda_ {0}$-$\Omega_{0}$ plane can be ruled out at a high confidence level) can be investigated. \subsection{The range of $\Omega_{0}$ and $\lambda_{0}$} The mass clustered with galaxies on smaller scales, $\Omega_{0,\mathrm{gal}}$, is 0.1 within a factor of two \citep[e.g.][]{PPeebles93a}. This lower limit is small compared to our full $\Omega_{0}$ range so we do not assume any prior lower limit on $\Omega_{0}$ except, of course, \begin{equation} \Omega_{0} \ge 0. \end{equation} Especially for comparison with other work it is important to note that, within the framework of cosmological models based on general relativity with which we (and almost everyone else at present) are working, $\Omega_{0} \ge 0$ is a \emph{requirement}. Results reported which include $\Omega_{0} < 0$ within the errors, or even as a best-fit value, do not indicate `implausible results' but merely improper statistics. Often, confidence contours are assumed to be ellipses and these are extended, if applicable, to $\Omega_{0} < 0$. (Of course, it is possible that $\Omega_{0} = 0$ is within the errors or even the best fit value for a certain set of results.) An extremely conservative upper limit comes from dynamical tests on larger (though still cosmologically small) scales; when this work was started, we assumed an (again, extremely conservative) upper limit $\Omega_0\le2$ \citep{OCzoske95a}. Since then, these methods have started to indicate smaller values of $\Omega_{0}$, \citep[e.g.][]{LdCostaNFGHSW97a} more in line with both a long tradition of low $\Omega_{0}$ values \citep[e.g.][]{JGottGST74a,PColesGEllis94a,PColesGEllis97a} (albeit with somewhat larger errors) as well as new determinations (often with quite small errors), examples of which are mentioned in Sect.~\ref{olprior}. We have assumed no prior upper or lower limits on $\lambda_{0}$ per se. This has two reasons: \begin{itemize} \item `Direct' measurements of $\lambda_{0}$ (as opposed to measurements of a combination of parameters involving $\lambda_{0}$) are virtually nonexistent. \item We obtain a small enough range in $\lambda_{0}$ from the values obtained from joint constraints on the range of $\Omega_{0}$ and $\lambda_{0}$. \end{itemize} Historically, positive $\lambda_{0}$ values have been considered more than negative ones, probably because positive values can have a wide range of relatively easily observable effects, while negative ones are more difficult to measure. Many cosmological tests have a degeneracy such that $\lambda_{0}$ and $\Omega_{0}$ are correlated, so that increasing $\lambda_{0}$ can be compensated for in some sense by increasing $\Omega_{0}$ as well. Thus, effects of negative values of $\lambda_{0}$ for a given value of $\Omega_{0}$ are hard to differentiate from the effects of larger values of $\Omega_{0}$ for larger (less negative) values of $\lambda_{0}$ or even $\lambda_{0}=0$. Here, we consider negative values of $\lambda_{0}$ as well. There is no a priori reason why they cannot exist. \emph{If} one believes that the `source' of $\lambda_{0}$ are zero-point fluctuations of a quantum vacuum, this would lend support to the idea that $\lambda_{0}>0$. However, it is not clear that this \emph{must} be the \emph{only} source of $\lambda_{0}$, and indeed it has been argued that, if this source of $\lambda_{0}$ exists, there must be an additional contribution with a \emph{negative} value \citep[e.g.][though the assumption that this is possible is so obvious to the authors it is barely stated!]{HMartelSW98a}. In spatially closed ($k=+1$) models, the antipode is required to be at $z > 4.5 $, the redshift of the most redshifted multiply imaged object currently known \citep{JGottPL89a,MParkJGott97a}.\footnote{Recently, a lensed object of even larger redshift has been detected at $z=4.92$ \citep{MFranxvDIKT98a}. However, at our resolution this would make only a negligible difference to the results so we have not updated the calculations to reflect this.} The light grey shaded area in the panel in the middle of the left column of Fig.~\ref{fi:prior} marks the right side of the region thus enclosed. This gives us a slightly $\Omega_{0}$-dependent upper limit on $\lambda_{0}$ which is slightly stronger than that obtained by merely excluding models with no big bang. (This can be done because these models have a maximum redshift which is less than the redshift of high-redshift objects, the only exception being some cosmological models which have $\Omega_{0} < 0.05$, the robust lower limit discussed above \citep[e.g.][]{BFeige92a}.) The age of the universe in units of the Hubble time, $H_{0}^{-1}$, is \begin{equation} \tau_0 = \int\limits_0^\infty \frac{\mathrm{d}z}{(1+z)\sqrt{Q(z)}} \label{eq:tau}, \end{equation} where $Q(z)$ is given by Eq.~(\ref{eq:Q}) and thus depends on $\Omega_{0}$ and $\lambda_{0}$. (There are world models in which the maximum redshift is not infinite but these are all models without a big bang and are excluded by the constraint from the antipodal redshift or the lower limit on $\Omega_{0}$ as discussed above and are thus not relevant for this work.) Clearly, in any physically realistic world model, $\tau_0H_{0}^{-1}$ exceeds the age of the oldest galactic globular clusters: \begin{equation} \tau_0 > t_{\mathrm{gc}}H_0. \label{eq:ineq} \end{equation} Following \citet{SCarrollPT92a}, we take a robust lower limit on $\lambda_{0}$ from conservative lower limits on the Hubble constant and age of the universe. This gives a lower limit on $\lambda_{0}$ from the value at $\Omega_{0} = 0$; at larger values of $\Omega_{0}$ the constraint on $\lambda_{0}$ is not as strict---by assuming the lower limit of $\lambda_{0} = -5$ independent of $\Omega_{0}$ we are being conservative. We choose $\lambda_{0}\ge -5$ instead of $\lambda_{0}\ge -7$ as in \citet{SCarrollPT92a} since no published current constraints examine this region in detail. (Were this the case, then including this area would be helpful if only to aid a direct comparison.) This value corresponds roughly to the \emph{one-sided} 99\% confidence level in the top row of Fig.~\ref{fi:prior} (see Sect.~\ref{olprior}), which is also a reason not to extend the area to more negative $\lambda_{0}$ values. \subsection{Prior probability for $\lambda_{0}$ and $\Omega_{0}$} \label{olprior} We have assumed no prior knowledge of $\lambda_{0}$ per se, apart from the upper and lower limits discussed above. This has three reasons: \begin{itemize} \item `Direct' measurements of $\lambda_{0}$ (as opposed to measurements of a combination of parameters involving $\lambda_{0}$) are virtually nonexistent. \item Based on general knowledge from the literature and our own low-resolution calculations, we expect lens statistics itself to constrain $\lambda_{0}$ quite well. \item Although recent measurements are encouraging (see Sect.~\ref{results}), the value of $\lambda_{0}$ is observationally not as well established as that of $\Omega_{0}$. \end{itemize} Regarding $t_{\mathrm{gc}}$ and $H_0$ as independent random quantities with known prior probability density functions $p(t_{\mathrm{gc}})$ and $p(H_0)$, the probability that Eq.~(\ref{eq:ineq}) is satisfied is \begin{equation} P(\tau_0 > t_{\mathrm{gc}}H_0) = \int\limits_0^\infty p(H_0) \int\limits_0^{\tau_0/H_0} p(t_{\mathrm{gc}}) \,\mathrm{d}t_{\mathrm{gc}}\,\mathrm{d}H_0. \label{eq:P} \end{equation} A cosmological world model is compatible with the absolute age of the oldest galactic globular clusters as long as the above expression does not vanish. Reasonably, we assume a prior probability density function that is proportional to this expression \begin{equation} p_1(\lambda_0,\Omega_0) = 1 \otimes \int\limits_0^\infty p(H_0) \int\limits_0^{\tau_0/H_0} p(t_{\mathrm{gc}}) \,\mathrm{d}t_{\mathrm{gc}}\,\mathrm{d}H_0. \label{eq:p1} \end{equation} The best estimate of the absolute age of the oldest galactic globular clusters currently is $t_{\mathrm{gc}}=11.5\pm1.3\,\mbox{Gyr}$ \citep{BChaboyerDKK98a}. We choose to formulate this prior information in the form of a lognormal distribution that meets these statistics \begin{equation} p(t_{\mathrm{gc}}) = L(t_{\mathrm{gc}}|11.5\,\mbox{Gyr}, 1.3\,\mbox{Gyr}). \label{eq:ptgc} \end{equation} Similarly, we roughly estimate $H_0=65\pm10\,\mbox{km}\,\mbox{s}^{-1}\,\mbox{Mpc}^{-1}$ and choose to formulate this prior information in form of a normal distribution \begin{equation} p(H_0) = N(H_0|65\,\mbox{km}\,\mbox{s}^{-1}\,\mbox{Mpc}^{-1}, 10\,\mbox{km}\,\mbox{s}^{-1}\,\mbox{Mpc}^{-1}), \label{eq:ph0} \end{equation} where the notation for $L$ and $N$ is such that the two arguments correspond to the mean and standard deviation. This estimate is compatible with `small' values of the Hubble constant, which is conservative in the sense that it restricts our region of the $\lambda_{0}$-$\Omega_{0}$ plane less than would `large' values. By the same token we neglect any time between the big bang and the formation of the oldest globular clusters. Inserting Eq.~(\ref{eq:ptgc}) and Eq.~(\ref{eq:ph0}) in Eq.~(\ref{eq:p1}) one obtains a well-founded a priori probability distribution for the parameters $\Omega_0$ und $\lambda_0$. Although observational evidence has always indicated a low value of $\Omega_{0}$ \citep[e.g.][]{JGottGST74a,PColesGEllis94a,PColesGEllis97a}, the inflationary paradigm \citep[e.g.][]{AGuth81a}, coupled with a prejudice against a non-negligible value of $\lambda_{0}$, has created a prejudice in favour of $\Omega_{0} = 1$,\footnote{After this was found to conflict with too many observations, the prejudice against a non-negligible value of $\lambda_{0}$ weakened, and the new prejudice has been in favour of a flat universe with $\lambda_{0} + \Omega_{0} = 1$.} unfortunately too often to the extent where this prior belief has been elevated to the status of dogma \citep[see, e.g.,][for an illuminating account]{DMatraversES95a} even though there are serious fundamental problems with the inflationary idea \citep[e.g.][]{RPenrose89a} and even though there might be other solutions to the problems it claims to solve \citep[e.g.][]{JBarrow95a,GCollins97a}. What is more, some current inflationary thinking \citep[e.g.][]{NTurokSHawking98a} seems able to predict values for $\lambda_{0}$ and $\Omega_{0}$ similar to current observationally determined values, though it would have been more interesting had this prediction been made before the recent improvements in the observational situation. (To be fair, many leading practitioners of inflation consider a flat universe to be a robust prediction and its observational falsification essentially a falsification of the entire paradigm.) Recently, in the light of overwhelming observational evidence in favour of a low value of $\Omega_{0}$ \citep[e.g.][]{RCarlbergetal97a,RCarlberg98a,RCarlbergetal98b, NBahcall97a,NBahcallFC97a,XFanBC97a,MBartelmannHCJP98a, CLineweaver98a}, whether determined more or less independently or in combination with other parameters, this prejudice is starting to weaken. Conservatively, these results can be summarised as \begin{equation} p_2(\lambda_0,\Omega_0) = L(\Omega_0|0.4,0.2). \label{eq:p2} \end{equation} A prior constraint on $\Omega_{0}$ is useful since lensing statistics alone, as expected and as our results show, cannot usefully constrain $\Omega_{0}$. In addition, we also consider the product of $p_2(\lambda_0,\Omega_0)$ with the age constraint $p_1(\lambda_0,\Omega_0)$, \begin{equation} p_3(\lambda_0,\Omega_0) = p_1(\lambda_0,\Omega_0) \otimes p_2(\lambda_0,\Omega_0). \label{eq:p3} \end{equation} \subsection{General discussion of prior information} Using harsher constraints would mean that results would reflect almost exclusively the prior information as opposed to the information derived from lensing statistics. It is not the purpose of this paper to do a joint analysis of several cosmological tests,\footnote{but see Sect.~\ref{results}} but rather to examine lens statistics as a cosmological test. For practical reasons, an upper limit on $\Omega_{0}$ and upper and lower limits on $\lambda_{0}$ are required. On the other hand, it is sensible to combine the results with conservative constraints from other well-understood cosmological tests where there is general agreement and little room for debate. Within our upper and lower limits, we present our results both with and without the constraints discussed above. The density values and confidence contours of the three prior probability density functions are shown in the right column of Fig.~\ref{fi:prior}. \begin{figure*} \hfill \resizebox{0.375\textwidth}{!}{\includegraphics{prior1.eps}} \noindent \resizebox{0.375\textwidth}{!}{\includegraphics{orient.eps}} \hfill \resizebox{0.375\textwidth}{!}{\includegraphics{prior2.eps}} \noindent \hfill \resizebox{0.375\textwidth}{!}{\includegraphics{prior3.eps}} \caption[]{ \emph{Left column:} The cosmological parameter plane. The four curved lines in the plot at the left are the isochrones \mbox{$t_0H_0=0.5,\dots,0.8$}. The straight line marks spatially flat world models. In the white region, the antipodal redshift falls below $z=4.5$, the redshift of the most redshifted multiply imaged object currently known \citep{JGottPL89a,MParkJGott97a}. \emph{Right column:} The prior probability distributions $p_1(\lambda_0,\Omega_0)$ (top panel), $p_2(\lambda_0,\Omega_0)$ (middle panel) and $p_3(\lambda_0,\Omega_0)$ (bottom panel). The pixel grey level is directly proportional to the probability density ratio, darker pixels reflect higher ratios. The pixel size reflects the resolution of our numerical computations. The contours mark 0.61, 0.26, 0.14 and 0.036 of the peak likelihood for the parameters $\lambda_0$ and~$\Omega_0$, which would correspond to the boundaries of the minimum $0.68$, $0.90$, $0.95$ and $0.99$ confidence regions \emph{if the distribution were Gaussian}} \label{fi:prior} \end{figure*} \section{Calculations} \label{calculations} Following K96, we use the \citet{GHinshawLKrauss87a} softened isothermal sphere model for modeling the light deflection properties of the lens galaxies. For this model, the lens equation reads \begin{equation} \label{eq:lens} x - y = \frac{bx}{\hat{s} + \sqrt{x^2 + \hat{s}^2}}, \end{equation} where $x$ is the angular position in the lens plane, $y$ the angular position in the source plane, $b\equiv4\pi(\sigma/c)^2(D_{\mathrm{ds}}/D_{\mathrm{os}})$, $\sigma$ denotes the one--dimensional velocity dispersion of the dark matter, $s$ denotes the core radius, $\hat{s}\equiv s/D_{\mathrm{od}}$ is the angular core radius and $D_{\mathrm{od}}$, $D_{\mathrm{os}}$ and $D_{\mathrm{ds}}$ denote the angular size distances between the observer and the lens galaxy, the observer and the source and the lens galaxy and the source, respectively. Still following K96, we model the distribution of elliptical and lenticular lens galaxies using Schechter functions with constant comoving density \begin{equation} n_{\mathrm{e}}=0.61\pm0.21\,h^3\,10^{-2}\,\mbox{Mpc}^{-3} \end{equation} ($h=H_0\,10^{-2}\,\mbox{km}^{-1}\,\mbox{s}\,\mbox{Mpc}$) and slope \begin{equation} \alpha_{\mathrm{e}}=-1.0\pm0.15. \end{equation} The lens galaxy luminosities are converted to the dark matter velocity dispersions of the softened isothermal lens model by means of Faber--Jackson type relations, \begin{equation} L/L_{*\mathrm{e}}=(\sigma/\sigma_{*\mathrm{e}})^{\gamma_{\mathrm{e}}}, \end{equation} where \begin{equation} \gamma_{\mathrm{e}}=4.0\pm0.5 \end{equation} and \begin{equation} \sigma_{*\mathrm{e}}=225.0\pm22.5\,\mathrm{km\,s}^{-1}. \end{equation} The core radii of the softened isothermal lens model are varied with the dark matter velocity dispersions according to \begin{equation} s/s_{*\mathrm{e}}=(\sigma/\sigma_{*\mathrm{e}})^{2+\varepsilon}, \end{equation} where $\varepsilon=2.8$ and $s_{*\mathrm{e}}=10h^{-1}\,\mbox{pc}$. We consider elliptical and lenticular lens galaxies only. For the number--magnitude counts of quasars, we adopt the best-fit model from K96. We neglect here evolution, dust and other possible systematic effects and refer the reader to K96 for a discussion. In our first calculations we apply Eq.~(\ref{eq:simplemarg}) and compute the a priori likelihood \begin{equation} p(D|\lambda_0,\Omega_0,\vec{\xi}_0) \end{equation} and the posterior probability density functions \begin{equation} p_1(\lambda_0,\Omega_0|D) = p(D|\lambda_0,\Omega_0,\vec{\xi}_0) \otimes p_1(\lambda_0,\Omega_0), \label{eq:simplemarg1} \end{equation} \begin{equation} p_2(\lambda_0,\Omega_0|D) = p(D|\lambda_0,\Omega_0,\vec{\xi}_0) \otimes p_2(\lambda_0,\Omega_0), \label{eq:simplemarg2} \end{equation} and \begin{equation} p_3(\lambda_0,\Omega_0|D) = p(D|\lambda_0,\Omega_0,\vec{\xi}_0) \otimes p_3(\lambda_0,\Omega_0) \label{eq:simplemarg3} \end{equation} in the limit where all nuisance parameters take precisely their mean values. To obtain an impression of the consequences of neglecting the uncerntainties of the nuisance parameters, in our second calculation we increase the value of the most uncertain nuisance parameter, $n_{\mathrm{e}}$, by two standard deviations. For the computation of the innermost integral on the right side of Eq.~(\ref{eq:p}), we consider the detectability of images in pairs: If the separation between the two closest images -- these are always images 2 and 3, counting from the outside in -- is more than the lower limit of the survey resolution limit $S(y)$, we define the image separation and flux ratio for the purpose of sample selection based on the two brightest images, usually 1 and 2. Otherwise we construct one image from the combined fluxes and flux-weighted positions of images 2 and 3 and define the image separation and flux ratio for the purpose of sample selection based on this combination image and image~1. In general, if the separation between images 1 and 2 is too large for the survey \emph{and} the separation between images 2 and 3 is large enough, then the image separation and flux ratio for the purpose of sample selection should be based on images 2 and 3. However, the present surveys are sensitive to the largest separations due to isolated galaxies, so this case doesn't need to be addressed in this paper (i.e.~implementing it would lead to the same results in the present case). For the calculation of the probabilities $p(m_i,z_i,\theta_i)$ the function $C(y)$ selects only those image configurations whose separation is $\pm10$ per cent of the observed separation $\theta_i$. Each of the three integrals on the right side of Eq.~(\ref{eq:p}) is approximated to an accuracy better than $0.004$ by a family of recursive monotone stable formulae \citep{PFavatiLR91a,PFavatiLR91b}. \section{Results and discussion} \label{results} \subsection{Information content} Given some observational data $D$, some model parameters $\vec{\phi}$, and some prior and posterior probability density functions $p(\vec{\phi})$ and $p(\vec{\phi}|D)$, the amount of information obtained from the data \citep[e.g.][]{JBernardoASmith94a} (on a logarithmic scale) is \begin{equation} \log[I(D)] = \int\limits p(\vec{\phi}) \log \left[ \frac{p(\vec{\phi}|D)}{p(\vec{\phi})} \right] \,\mathrm{d}\vec{\phi}. \label{eq:info} \end{equation} The amounts of information obtained from our sample data are given in the caption of Fig.~\ref{fi:posterior}. \subsection{Results} \begin{figure*} \resizebox{0.375\textwidth}{!}{\includegraphics{odlik.eps}} \hfill \resizebox{0.375\textwidth}{!}{\includegraphics{oxlik.eps}} \caption[]{ \emph{Left panel:} The likelihood function $p(D|\lambda_0,\Omega_0,\vec{\xi}_0)$. All nuisance parameters are assumed to take precisely their mean values. The pixel grey level is directly proportional to the likelihood ratio, darker pixels reflect higher ratios. The pixel size reflects the resolution of our numerical computations. The contours mark the boundaries of the minimum $0.68$, $0.90$, $0.95$ and $0.99$ confidence regions for the parameters $\lambda_0$ and $\Omega_0$. \emph{Right panel:} Exactly the same as the left panel, but the parameter $n_{\mathrm{e}}$ is increased by two standard deviations} \label{fi:likelihood} \end{figure*} The left panel of Fig.~\ref{fi:likelihood} shows the constraints on the cosmological parameters $\lambda_{0}$ and $\Omega_{0}$ based only on the information obtained from the lens statistics. Quite good constraints can be placed on $\lambda_{0}$, more or less independent of $\Omega_{0}$. It is a well-known fact (see K96 and references therein) that lensing statistics can provide a good \emph{upper} limit on $\lambda_{0}$. While in the past this has mainly been discussed in the context of flat cosmological models, it is of course more general \citep{SCarrollPT92a,EFalcoKM98a}. Although no unexpected effects are seen, it is important to note that this is the first time $\lambda_{0}$ and $\Omega_{0}$ have been used as independent parameters in conjunction with a non-singular lens model in an analysis of this type. Our analysis shows for the first time that gravitational lensing statistics can place a quite firm \emph{lower} limit on $\lambda_{0}$ as well, again more or less independent of $\Omega_{0}$. The constraint is not as tight since the gradient in the probability density is not as steep towards negative $\lambda_{0}$ as towards positive $\lambda_{0}$. If this lower limit can be improved enough, it could provide an independent confirmation of the detection of a positive cosmological constant (see Sect.~\ref{compare}). On the other hand, this might be difficult, since Poisson errors in the number of lenses and uncertainties in the normalisation of the luminosity density of galaxies introduce relatively large uncertainties in this region of parameter space \citep[K96,][]{EFalcoKM98a}. The latter effect is illustrated in the right panel of Fig.~\ref{fi:likelihood}, where $n_{\mathrm{e}}$, the galaxy luminosity density normalisation, is increased by two standard deviations: the derived lower limit on $\lambda_{0}$ changes much more than does the upper limit. Nevertheless, our robust lower limit is much better than the $-7$ mentioned in \citet{SCarrollPT92a}. Our results place no useful constraints on $\Omega_{0}$. It is interesting to note the fact, however, that likely values of $\lambda_{0}$ and $\Omega_{0}$ are positively correlated. This is similar to most cosmological tests, a notable exception being constraints derived from CMB anisotropies (see Sect.~\ref{compare}). Fortunately, constraints on $\Omega_{0}$ from other sources are quite good (Sect.~\ref{olprior}). Often, this is cast in the form of a constraint on $\Omega_{0} - \lambda_{0}$ \citep[e.g.][]{ACoorayQC99a} or, perhaps more practical, $\lambda_{0} - \Omega_{0}$. This is a reasonable way or reducing the information to one number, at least when one is concerned with upper limits on $\lambda_{0}$ (or $\lambda_{0} - \Omega_{0}$) in a relatively low-density universe. Besides the obvious dependencies on confidence levels and assumptions made, when comparing constraints on $\lambda_{0}$ from different investigations one should keep in mind whether they are approximations, like $\lambda_{0} - \Omega_{0}$ in lensing statistics, and whether a value for a particular scenario (for example, for a flat universe) is the `obvious' definition or in fact describes the intersection of the $k=0$ line with the corresponding 2-dimensional confidence contour, which in general will give a different number. Also, some authors plot `real' confidence contours while some actually plot contours at values which would correspond to certain confidence contours were the likelihood distribution in the parameter space in question Gaussian. \begin{figure*} \resizebox{0.375\textwidth}{!}{\includegraphics{odpost1.eps}} \hfill \resizebox{0.375\textwidth}{!}{\includegraphics{oxpost1.eps}} \noindent \resizebox{0.375\textwidth}{!}{\includegraphics{odpost2.eps}} \hfill \resizebox{0.375\textwidth}{!}{\includegraphics{oxpost2.eps}} \noindent \resizebox{0.375\textwidth}{!}{\includegraphics{odpost3.eps}} \hfill \resizebox{0.375\textwidth}{!}{\includegraphics{oxpost3.eps}} \caption[]{ \emph{Left column:} The posterior probability density functions $p_1(\lambda_0,\Omega_0|D)$ (top panel), $p_2(\lambda_0,\Omega_0|D)$ (middle panel) and $p_3(\lambda_0,\Omega_0|D)$ (bottom panel). All nuisance parameters are assumed to take precisely their mean values. The pixel grey level is directly proportional to the likelihood ratio, darker pixels reflect higher ratios. The pixel size reflects the resolution of our numerical computations. The contours mark the boundaries of the minimum $0.68$, $0.90$, $0.95$ and $0.99$ confidence regions for the parameters $\lambda_0$ and $\Omega_0$. The respective amounts of information (Eq.~(\ref{eq:info})) obtained from our sample data are $I_1=1.74$, $I_2=1.24$ and $I_3=1.74$. \emph{Right column:} Exactly the same as the left column, but the parameter $n_{\mathrm{e}}$ is increased by two standard deviations} \label{fi:posterior} \end{figure*} The left plot in the top row of Fig.~\ref{fi:posterior} shows the joint likelihood of our lensing statistics analysis and that obtained by using conservative estimates for $H_{0}$ and the age of the universe (see Sect.~\ref{olprior}). Although neither method alone sets useful constraints on $\Omega_{0}$, their combination does, since the constraint involving $H_{0}$ and the age of the universe only allows large values of $\Omega_{0}$ for $\lambda_{0}$ values which are excluded by lens statistics. Even though the 68\%~contour still allows almost the entire $\Omega_{0}$ range, it is obvious from the grey scale that much lower values of $\Omega_{0}$ are favoured by the joint constraints. The upper limit on $\lambda_{0}$ changes only slightly while, as is to be expected, the lower limit becomes tighter. Also, the change caused by increasing $n_{\mathrm{e}}$ by 2 standard deviations is less pronounced, with regard to both lower and upper limits on $\lambda_{0}$, as demonstrated in the right plot in top row of Fig.~\ref{fi:posterior}. The middle row of Fig.~\ref{fi:posterior} shows the effect of including our prior information on $\Omega_{0}$ (see Sect.~\ref{olprior}). As is to be expected, (for both values of $n_{\mathrm{e}}$) lower values of $\Omega_{0}$ are favoured. This has the side effect of weakening our lower limit on $\lambda_{0}$ (though only slightly affecting the upper limit). We believe that the left plot of the bottom row of Fig.~\ref{fi:posterior} represents very robust constraints in the $\lambda_{0}$-$\Omega_{0}$ plane. The upper limits on $\lambda_{0}$ come from gravitational lensing statistics, which, due to the extremely rapid increase in the optical depth for larger values of $\lambda_{0}$, are quite robust and relatively insensitive to uncertainties in the input data (compare the left and right columns of Fig.~\ref{fi:posterior}) as well as to the prior information used data (compare the upper, lower and middle rows of Fig.~\ref{fi:posterior}). The upper and lower limits on $\Omega_{0}$ are based on a number of different methods and appear to be quite robust, as discussed in Sect.~\ref{olprior}. The combination of the relatively secure knowledge of $H_{0}$ and the age of the universe combine with lens statistics to produce a good lower limit on $\lambda_{0}$, although this is to some extent still subject to the caveats mentioned above. If one is interested in the allowed range of $\lambda_{0}$, one can marginalise over $\Omega_{0}$ to obtain a probability distribution for $\lambda_{0}$. This is illustrated in Fig.~\ref{fi:marginal} \begin{figure*} \resizebox{0.375\textwidth}{!}{\includegraphics{odmar.eps}} \hfill \resizebox{0.375\textwidth}{!}{\includegraphics{oxmar.eps}} \noindent \resizebox{0.375\textwidth}{!}{\includegraphics{odcum.eps}} \hfill \resizebox{0.375\textwidth}{!}{\includegraphics{oxcum.eps}} \hfill \caption[]{ \emph{Left column:} The top panel shows the normalised marginal likelihood function $p(\lambda_0|D)$ (light grey curve) and the marginal posterior probability density functions $p_1(D|\lambda_0)$ (medium grey curve), $p_2(D|\lambda_0)$ (dark grey curve) and $p_3(D|\lambda_0)$ (black curve). All nuisance parameters are assumed to take precisely their mean values. The bottom panel shows the respective cumulative distribution functions. These can be used to construct any desired $\Omega_{0}$-averaged upper or lower limits on $\lambda_{0}$. \emph{Right column:} Exactly the same as the left column, but the parameter $n_{\mathrm{e}}$ is increased by two standard deviations} \label{fi:marginal}% \end{figure*} and Table~\ref{ta:results}. \begin{table*} \caption[]{ Marginal mean values, standard deviations and $0.95$ confidence intervals for the parameter~$\lambda_0$ on the basis of the marginal distributions shown in the top row of Fig.~\ref{fi:marginal}; `information' refers to Eq.~\ref{eq:info}} \label{ta:results} \begin{tabular*}{\linewidth}{@{\extracolsep{\fill}}lccccc} \hline \hline Distribution & Mean & standard deviation & \multicolumn{2}{l}{95\% c.l.~range} & information \\ \hline $p(D|\lambda_0)$ & $-0.35$ & $1.07$ & $-2.55$ & $1.51$ & \\ $p_1(\lambda_0|D)$ & $-0.02$ & $0.80$ & $-1.59$ & $1.50$ & $1.74$ \\ $p_2(\lambda_0|D)$ & $-0.78$ & $0.97$ & $-2.85$ & $0.76$ & $1.24$ \\ $p_3(\lambda_0|D)$ & $-0.34$ & $0.67$ & $-1.72$ & $0.79$ & $1.74$ \\ \hline \end{tabular*} \end{table*} \subsection{Comparison with other results} \label{compare} For comparison with other results, as a first step one can examine the allowed range of $\lambda_{0}$ for the current `best-fit' value for $\Omega_{0}$, which we take, based on the work cited in Sect.~\ref{olprior}, to be $\Omega_{0}=0.3$. (A more conservative estimate is reflected by using the prior probability distribution $p_2(\lambda_0,\Omega_0) = L(\Omega_0|0.4,0.2)$ as shown by the dark grey curve in Fig.~\ref{fi:marginal} and in Table~\ref{ta:results}.) On the other hand, previous limits on $\lambda_{0}$ have often been quoted for a flat universe (K96 and references therein). We consider both cases in Tables~\ref{ta:specialo} and \ref{ta:specialk}. \begin{table*} \begin{minipage}[t]{\textwidth} \caption[]{ Mean values and ranges for assorted confidence levels for the parameter $\lambda_{0}$ for our a priori and various a posteriori likelihoods from this analysis and from other tests from the literature (using the latest publicly available results) for the special case $\Omega_{0} = 0.3$. Except where noted, the ranges quoted are the projections of the corresponding confidence contours in the $\lambda_{0}$-$\Omega_{0}$ plane onto the $\lambda_{0}$ axis\footnote{Note that some references quote confidence ranges for $k = 0$---in general, these will be different than the projection of the intersection of the corresponding contour in the $\lambda_{0}$-$\Omega_{0}$ plane onto the $\lambda_{0}$-axis.} (as opposed to $\Omega_{0}$-independent estimates, which of course would always give a smaller range), and are of course two-sided, not one-sided, bounds. Values are either those quoted in the references given and/or obtained from figures in those references; inequalities mean that the corresponding confidence contour is to be found in the range indicated by the inequality, e.g.~$<-1.2$ would mean that the corresponding contour level is to be found at $\lambda_{0}<-1.2$, \emph{not} that the constraint is $\lambda_{0}<-1.2$ at the corresponding confidence level. This arises because the corresponding area of parameter space was not examined in the reference in question. If the confidence interval could not be determined from the reference, \emph{both} values in the corresponding column are missing} \label{ta:specialo} \begin{tabular*}{\linewidth}{@{\extracolsep{\fill}}lrrrrrrrr} \hline \hline Cosmological test& \multicolumn{2}{c}{68\% c.l.~range} & \multicolumn{2}{c}{90\% c.l.~range} & \multicolumn{2}{c}{95\% c.l.~range} & \multicolumn{2}{c}{99\% c.l.~range} \\ \hline this work, $p(D|\lambda_0)$ & $-1.18$ & $0.24$ & $-2.19$ & $0.50$ & $-2.81$ & $0.60$ & $-4.16$ & $0.73$ \\ this work, $p_1(\lambda_0|D)$ & $-0.97$ & $0.46$ & $-1.55$ & $0.60$ & $-1.89$ & $0.69$ & $-2.73$ & $0.81$ \\ this work, $p_2(\lambda_0|D)$ & $-2.00$ & $0.49$ & $-3.33$ & $0.65$ & $-4.10$ & $0.72$ & $<-5.00$ & $0.80$ \\ this work, $p_3(\lambda_0|D)$ & $-1.20$ & $0.52$ & $-1.98$ & $0.69$ & $-2.35$ & $0.77$ & $-3.40$ & $0.86$ \\ lens statistics (K96) & \multicolumn{8}{c}{not possible since only $ k=0 $ models considered} \\ radio lenses\footnote{\citet{EFalcoKM98a}}\footnote{contour at 95.4\% not 95\%} & $-0.54$ & $0.28$ & $<-1.00$ & $0.75$ & $<-1.00$ & $0.89$& $$ & $$ \\ optical lenses\footnote{\citet{EFalcoKM98a}}\footnote{contour at 95.4\% not 95\%} & $<-1.00$ & $0.37$ & $<-1.00$ & $0.75$ & $<-1.00$ & $0.89$ & $$ & $$ \\ radio + optical lenses\footnote{\citet{EFalcoKM98a}}\footnote{contour at 95.4\% not 95\%} & $<-1.00$ & $-0.12$ & $<-1.00$ & $0.50$ & $<-1.00$ & $0.70$ & $<-1.00$ & $0.89$ \\ supernovae $m$-$z$ relation $\cal A$\footnote{\citet{SPerlmutteretal98a}} & $-0.70$ & $0.50$ & $-1.15$ & $0.75$ & $$ & $$ & $$ & $$ \\ supernovae $m$-$z$ relation $\cal B$\footnote{\citet{ARiessetal98a}}\footnote{Fig.~6, solid contours}\footnote{contours at 68.3\%, 95.4\% and 99.7\% instead of 68\%, 95\% and 99\% respectively} & $0.78$ & $1.00$ & $$ & $$ & $0.53$ & $1.27$ & $0.27$ & $1.41$ \\ CNOC survey\footnote{\citet{RCarlberg98a}} & $<-0.50$ & $<-0.50$ & $<-0.50$ & $<-0.50$ & $$ & $$ & $$ & $$ \\ CMB\footnote{\citet{CLineweaver98a}}\footnote{contours at 68.3\%, 95.4\% and 99.7\% instead of 68\%, 95\% and 99\%, respectively} & $0.44$ & $0.67$ & $$ & $$ & $0.36$ & $>0.90$ & $0.26$ & $>0.90$ \\ CMB + \textit{IRAS}\footnote{\citet{MWebsterBHLLR98a}} & \multicolumn{8}{c}{not possible since only $ k=0 $ models considered} \\ double radio sources\footnote{\citet{EGuerraDW98a}} & $0.00$ & $1.00$ & $<-2.00$ & $1.39$ & $$ & $$ & $$ & $$ \\ \hline \end{tabular*} \end{minipage} \end{table*} \begin{table*} \begin{minipage}[t]{\textwidth} \caption[]{ Mean values and ranges for assorted confidence levels for the parameter $\lambda_{0}$ for our a priori and various a posteriori likelihoods from this analysis and from other tests from the literature (using the latest publicly available results) for the special case $k = 0$. Otherwise the same as Table~\ref{ta:specialo}, in particular the references are not listed in the footnotes to this table. $X$ denotes the fact that there is no intersection of the confidence contour with the $k=0$ line} \label{ta:specialk} \begin{tabular*}{\linewidth}{@{\extracolsep{\fill}}lrrrrrrrr} \hline \hline Cosmological test& \multicolumn{2}{c}{68\% c.l.~range} & \multicolumn{2}{c}{90\% c.l.~range} & \multicolumn{2}{c}{95\% c.l.~range} & \multicolumn{2}{c}{99\% c.l.~range} \\ \hline this work, $p(D|\lambda_0)$ & $-0.68$ & $0.51$ & $<-1.00$ & $0.57$ & $<-1.00$ & $0.62$ & $<-1.00$ & $0.70$ \\ this work, $p_1(\lambda_0|D)$ & $-0.09$ & $0.56$ & $-0.38$ & $0.64$ & $-0.57$ & $0.68$ & $-1.04$ & $0.81$ \\ this work, $p_2(\lambda_0|D)$ & $X$ & $X$ & $0.09$ & $0.69$ & $-0.03$ & $0.73$ & $-0.28$ & $0.92$ \\ this work, $p_3(\lambda_0|D)$ & $0.47$ & $0.48$ & $0.18$ & $0.67$ & $0.07$ & $0.70$ & $-0.14$ & $0.84$ \\ lens statistics\footnote{value for $k=0$, not projection} & $$ & $$ & $$ & $$ & $<0.00$ & $0.66$ & $$ & $$ \\ radio lenses\footnote{contour at 95.4\% not 95\%} & $-0.47$ & $0.56$ & $<-1.00$ & $0.72$ & $<-1.00$ & $0.80$ & $<-1.00$ & $0.85$ \\ optical lenses\footnote{contour at 95.4\% not 95\%} & $<-1.00$ & $0.56$ & $<-1.00$ & $0.72$ & $<-1.00$ & $0.80$ & $<-1.00$ & $0.87$ \\ radio + optical lenses\footnote{contour at 95.4\% not 95\%} & $-0.87$ & $0.43$ & $<-1.00$ & $0.60$ & $<-1.00$ & $0.69$ & $<-1.00$ & $0.78$ \\ supernovae $m$-$z$ relation $\cal A$ & $0.20$ & $0.60$ & $-0.05$ & $0.75$ & $$ & $$ & $$ & $$ \\ supernovae $m$-$z$ relation $\cal B$\footnote{Fig.~6, solid contours}\footnote{contours at 68.3\%, 95.4\% and 99.7\% instead of 68\%, 95\% and 99\% respectively} & $0.74$& $0.83$ & $$ & $$ & $0.61$ & $0.92$ & $0.50$ & $1.00$ \\ CNOC survey & $0.85$ & $0.95$ & $0.81$ & $0.98$ & $$ & $$ & $$ & $$ \\ CMB\footnote{contour at 68.3\% instead of 68\%; other contours, and part of the 68.3\% contour, lie partially in the $k=+1$ area of parameter space which was not examined for technical reasons in \citet{CLineweaver98a}} & $<0.00$ & $0.60$ & $<0.00$ & $<0.00$ & $<0.00$ & $<0.00$ & $<0.00$ & $<0.00$ \\ CMB + \textit{IRAS}\footnote{value for $k=0$, not projection} & $0.47$ & $0.71$ & $$ & $$ & $$ & $$ & $$ & $$ \\ double radio sources & $0.35$ & $1.00$ & $0.70$ & $1.00$ & $$ & $$ & $$ & $$ \\ \hline \end{tabular*} \end{minipage} \end{table*} We do not do a comparison for the special case $\lambda_{0} = 0$ since this analysis of gravitational lensing statistics does not usefully constrain $\Omega_{0}$ (any limits coming only from the prior information on $\Omega_{0}$). It is beyond the scope of this paper to do a full comparison of different cosmological tests. Except for a few general comments, we therefore restrict ourselves to comments on the similarities and differences between the results from this work without using prior information on $\lambda_{0}$ and $\Omega_{0}$, i.e.~(the left plot in) Fig.~\ref{fi:likelihood}, and the those from K96 and \citet{EFalcoKM98a} (using only optical data, i.e.~the lower left plot in their Fig.~5). Taking all results at face value and examining the $\Omega_{0} = 0.3$ case first, we note that with `three-and-one-half' exceptions (counting as one test each the four from this work and the three from \citet{EFalcoKM98a}) the 68\% c.l.~\emph{lower} limit from \citet{CLineweaver98a} is \emph{higher} than \emph{all} 68\% \emph{upper} limits from other tests, while the 95\% lower and upper confidence levels from \citet{CLineweaver98a} are higher than the corresponding limits from the other tests for all but one of these. Even at the 99.9\% confidence level (not shown in Table~\ref{ta:specialo}), the \citet{CLineweaver98a} result requires $\lambda_{0} \ge 0.12$. If one assumes $\Omega_{0} = 0.3$, only \citet{CLineweaver98a} requires $\lambda_{0} >0$, though all other tests (except \citet{RCarlberg98a}) are compatible with this. This is not surprising, since it is well-known that constraints from CMB anisotropies tend to run more or less orthogonal in the $\lambda_{0}$-$\Omega_{0}$ plane to those from most other tests \citep[e.g.][]{MWhite98a,DEisensteinHT98b,MTegmarkEH98a,MTegmarkEHK98a}. Examining the $k=0$ case, it is interesting to note that the 68\% (90\%) confidence level \emph{lower} limit on $\lambda_{0}$ from \citet{RCarlberg98a} is \emph{higher} than \emph{all} of the 68\% (90\%) c.l.~\emph{upper} limits from \emph{all} other tests except \citet{EGuerraDW98a}. Otherwise, with `one-and-one-half' exceptions all tests are compatible even at the 68\% confidence level. If one assumes $k = 0$, then the evidence for $\lambda_{0} > 0$ looks convincing: at the 68\% confidence level, again with `one-and-one-half' exceptions, all tests indicate $\lambda_{0} > 0$; even at 90\% the evidence is still quite good, if one keeps in mind that the gradient towards smaller values of $\lambda_{0}$ is generally not as steep as towards larger values. Again taken at face value, neither the $k = 0$ case nor the $\Omega_{0} = 0.3$ case are compatible with all tests, even at the $\approx$90\% confidence level. It appears the simplest solution to achieve concordance would be to have $\Omega_{0} \approx 0.2$, which is within the error on $\Omega_{0}$ discussed in Sect.~\ref{olprior}. For $k=0$ this would imply $\lambda_{0} = 0.8$, which seems to be ruled out, thus ruling out the flat universe altogether. For a non-flat universe, reducing $\Omega_{0}$ would, due to the CMB constraint, require a higher value of $\lambda_{0}$, and thus make the $\lambda_{0} = 0$ case more unlikely, ruling out this special case as well. On balance, a cosmological model with $\lambda_{0} \approx 0.3$ and $\Omega_{0} \approx 0.25$ seems compatible with all known observational data (not just those discussed here) at a comfortable confidence level. For a `likely' $\Omega_{0}$ value of 0.3 we have calculated the likelihood with the higher resolution $\Delta\lambda_{0}=0.01$. This is shown in Fig.~\ref{fi:03}. \begin{figure*} \resizebox{0.375\textwidth}{!}{\includegraphics{odlik_03.eps}} \hfill \resizebox{0.375\textwidth}{!}{\includegraphics{odcum_03.eps}} \caption[]{ \emph{Left panel:} The likelihood function as a function of $\lambda_{0}$ for $\Omega_{0}=0.3$ and with all nuisance parameters taking their default values. \emph{Right panel:} The same but plotted cumulatively. See Table~\ref{ta:03}} \label{fi:03} \end{figure*} From these calculations one can extract confidence limits which, due to the higher resolution in $\lambda_{0}$, are more accurate. These are presented in Table~\ref{ta:03} and should be compared to those for $p(D|\lambda_0)$ from Table~\ref{ta:specialo}. \begin{table*} \caption[]{ Confidence ranges for $\lambda_{0}$ assuming $\Omega_{0}=0.3$. Unlike the results presented in Table~\ref{ta:specialo}, these figures are for a specific value of $\Omega_{0}$ and not the values of intersection of particular contours with the $\Omega_{0}=0.3$ line in the $\lambda_{0}$-$\Omega_{0}$ plane. These are more appropriate if one is convinced that $\Omega_{0}=0.3$ and have been calculated using ten times better resolution than the rest of our results presented in this work. See Fig.~\ref{fi:03}} \label{ta:03} \begin{tabular*}{\linewidth}{@{\extracolsep{\fill}}rrrrrrrr} \hline \hline \multicolumn{2}{c}{68\% c.l.~range} & \multicolumn{2}{c}{90\% c.l.~range} & \multicolumn{2}{c}{95\% c.l.~range} & \multicolumn{2}{c}{99\% c.l.~range} \\ \hline $-1.27$ & $0.27$ & $-2.26$ & $0.51$ & $-2.87$ & $0.60$ & $-4.10$ & $0.72$ \\ \hline \end{tabular*} \end{table*} Again, a full discussion of joint constraints involving discussion of possible sources of error for each test, as well as comparing the full contours in the $\lambda_{0}$-$\Omega_{0}$ plane, is beyond the scope of this paper. However, quick comparisons would be aided were the results of all tests available in an easy-to-process electronic form (see below); such quick consistency tests would enable one to spot areas of inconsistency much more quickly. Also, it should be emphasised that the projections onto the $\lambda_{0}$-axis of the intersection of a particular confidence contour with the $\Omega_{0}=0.3$ or $k=0$ axis are generally \emph{not} the same as the corresponding confidence interval for the $\Omega_{0}=0.3$ or $k=0$ special cases. For a flat universe, our 95\% confidence level upper limit on $\lambda_{0}$-$\Omega_{0}$, i.e.~the value of $\lambda_{0}$ where this contour crosses the $k = 0$ line, is $\lambda_{0} < 0.62$. This is essentially the same as the $\lambda_{0} < 0.66$ of K96, as was to be expected considering we used essentially the same data and methods. Interpreted cautiously, one might conclude from this that the singular isothermal sphere model is a good approximation as far as determining the cosmological parameters from lens statistics is concerned, as was assumed in \citet{EFalcoKM98a}. Our 99\% confidence level upper limit on $\lambda_{0}$ is $\lambda_{0} < 0.70$. This is quite a tight upper bound on $\lambda_{0}$ and appears to be quite robust. Perhaps more interesting is the comparison with (the results using only optical data in) \citet{EFalcoKM98a}. Although a detailed comparison is complicated by the different plotting scheme and reducing the entire contour (or indeed grey-scale) plot to a few numbers throws away information, it is obvious that the plots are broadly similar. Our 68\% contour is, for $\Omega_{0} \approx 1$, roughly parallel to the $\lambda_{0}$-axis at $\lambda_{0}\approx -1$. This is just at the edge of the \citet{EFalcoKM98a} plot, and as they provide no grey-scale, it is difficult to compare the lower limits on $\lambda_{0}$. Thus, while our main goal was to explore a `large enough' region of parameter space, comparison in the areas where there is overlap shows consistency, which strengthens our faith in the conclusions pertaining to areas of parameter space where there is no overlap. Recently, it has become quite fashionable to discuss joint constraints derived from a variety of cosmological tests. This has grown from plotting the overlap of likelihood contours (often in a space spanned by parameters other than $\lambda_{0}$ and $\Omega_{0}$) \citep[e.g.][]{JOstrikerPSteinhardt95a,MTurner96a,JBaglaPN96a,LKrauss98a, MWhite98a} to full-blown joint likelihood analyses, both detailed theoretical investigations of what will be possible in the future \citep[e.g.][]{MTegmarkEH98a,MTegmarkEHK98a,DEisensteinHT98a, DEisensteinHT98b} and more restricted analyses using present data \citep[e.g.][]{MWebsterBHLLR98a}. While in some cases it is quick and easy to calculate the likelihood as a function of $\lambda_{0}$ and $\Omega_{0}$ given the data, for example for tests using the $m$-$z$ relation, in other cases such as the present one it is a major programming and computational effort to do so. To aid comparisons, all figures from this paper are available in the form of tables of numbers at \begin{quote} \verb|http://multivac.jb.man.ac.uk:8000/ceres|\\ \verb|/data_from_papers/lower_limit.html| \end{quote} and we urge our colleagues to follow our example. We applaud the fact that most results are now presented in the $\lambda_{0}$-$\Omega_{0}$ plane, as opposed to using other parameters such as $q_{0}$ or $\Omega_{\mathrm{tot}}\equiv\lambda_{0}+\Omega_{0}$. A further aid in comparison would be a uniform choice of axes. We prefer to plot $\Omega_{0}$ on the $y$-axis and $\lambda_{0}$ on the $x$ axis since up/down symmetry is less fundamental than left/right symmetry and this mirrors the fact that $\Omega_{0}$ has the physical lower limit $\Omega_{0} = 0$ whereas no corresponding upper or lower limits for $\lambda_{0}$ exist. Square plots with the same range would further aid the comparison. Of course, if all data are publicly available, then they can be re-plotted to taste. \section{Summary and conclusions} \label{conclusions} We have re-analysed optical gravitational lens surveys from the literature, using the techniques described in \citet{CKochanek96a}, for the first time allowing both the cosmological constant $\lambda_{0}$ and the density parameter $\Omega_{0}$ to be free parameters while also using a non-singular lens model. We confirm the well-known results that gravitational lensing statistics can provide a good upper limit on $\lambda_{0}$ but are relatively insensitive to $\Omega_{0}$. We have presented the new result of a robust lower limit on $\lambda_{0}$, which is a substantial improvement on previously known \emph{robust} lower limits. Coupled with relatively conservative prior information about the Hubble constant $H_{0}$, the age of the universe and the well-established value of $\Omega_{0}$, one can reduce the allowed parameter space in the $\lambda_{0}$-$\Omega_{0}$ plane to a small, finite region, which is similar to the area allowed by joint constraints based on many other cosmological tests (see Fig.~\ref{fi:posterior}). Using lens statistics information alone, at 95\% confidence, our lower and upper limits on $\lambda_{0}-\Omega_{0}$ are respectively $-3.17$ and $0.3$. For a flat universe, this corresponds to lower and upper limits on $\lambda_{0}$ of respectively $-1.09$ and $0.65$. Keeping in mind the difficulties of a quantitative comparison, this is in good agreement with other recent measurements of the cosmological constant. This value was calculated from Table~\ref{ta:03} and assuming a degeneracy in $\lambda_{0} - \Omega_{0}$ as in \citet{ACoorayQC99a} and \citet{ACooray99a}. For comparison, from Table~\ref{ta:specialk}, the corresponding value for the upper limit on $\lambda_{0}$ is $0.62$ and the value from K96 is $0.66$.\footnote{The value from \citet{ACoorayQC99a} and \citet{ACooray99a} is 0.79, but it should be noted this value (the same in both papers) is based on different surveys, namely the Hubble Deep Field and CLASS, respecively.)} For detailed comparison of cosmological tests, one needs to compare confidence contours---calculated in the same, preferably in the `real', way---in the same parameter space. Of course, this makes it difficult to meaningfully reduce the results of a given cosmological test to one or even a few single numbers. Unless a cosmological test is developed which can measure $\lambda_{0}$ independently of any other parameters, there is not much point in quoting unqualified `limits on $\lambda_{0}$'. Presently tentative claims of the detection of a positive cosmological constant, if true, would rank among the great discoveries of cosmology. Even though there are serious difficulties involved, it seems worthwhile to be able to confirm this result by improving the lower limit on $\lambda_{0}$ derived from gravitational lensing statistics. Targetting the two primary sources of uncertainty calls for improving our knowledge of the normalisation of the local luminosity density of galaxies as well as increasing the size of gravitational lens surveys. As far as the latter goes, the CLASS survey \citep{IBrowneJAHMNWBKBFBPRWP97a,SMyersetal99a} looks the most promising at the moment. In a companion paper \citep{PHelbigMQWBK99a}, we have shown that comparable constraints to the ones presented in this work can be obtained from the JVAS gravitational lens survey; this gives us hope that the much larger CLASS survey will offer improvement in this area. Cosmological tests which set tight upper limits on $\Omega_{0}$ imply, for a flat $k = 0$ universe, a value of $\lambda_{0}$ which is ruled out by lensing statistics. For a non-flat universe, many tests are indicating $\lambda_{0} > 0$, and at present a cosmological model with $\lambda_{0} \approx 0.3$ and $\Omega_{0} \approx 0.25$ seems compatible with all known observational data, with neither a flat universe nor a universe without a positive cosmological constant being viable alternatives. The simplest case, the Einstein-de~Sitter universe with $\lambda_{0} = 0$ and $\Omega_{0} = 1$, both flat and without a cosmological constant, had been abandoned long before the new observational data cited in this work came to light \citep[see, e.g.,][and references therein]{JOstrikerPSteinhardt95a}; this trend has continued, with the next-most-simple cases also no longer viable. For $\lambda_{0}$ and $\Omega_{0}$, we have in a sense reached the least simple case; it will be interesting to see if this trend continues with regard to the other cosmological parameters, in particular those which can be measured by the \textit{Planck Surveyor} mission. Larger gravitational lens surveys such as CLASS will be a step in this direction. \begin{acknowledgements} We thank Sjur Refsdal, Leon Koopmans, Lutz Wisotzki and many colleagues at Jodrell Bank for helpful comments and suggestions. This research was supported in part by the European Commission, TMR Programme, Research Network Contract ERBFMRXCT96-0034 `CERES'. \end{acknowledgements}
\section{Introduction} Hybrid baryons are bound states of three quarks with an explicit excitation in the gluon field of QCD. The construction of (hybrid) baryons in a model motivated from and consistent with lattice gauge theory, the non--relativistic flux--tube model of Isgur and Paton, was detailed in ref. \cite{hadron} There we studied the detailed flux dynamics and built the flux hamiltonian. A minimal amount of quark motion is allowed in response to flux motion, in order to work in the centre of mass frame. Otherwise, we make the so--called ``adiabatic'' approximation, where the flux motion adjusts itself instantaneously to the motion of the quarks. The main result was that the lowest flux excitation can to a high degree of accuracy (about 5\%) be simulated by neglecting all flux--tube motions except the vibration of a junction. The junction acquires an effective mass from the motion of the remainder of the flux--tube and the quarks. The model is then simple: a junction is connected via a linear potential to the three quarks. The ground state of the junction motion corresponds to a conventional baryon and the various excited states to hybrid baryons. \section{Quantum numbers of (hybrid) baryons} \begin{table}[t] \begin{center} \caption{\small Quantum numbers of ground state (hybrid) baryons for the adiabatic surfaces $B,H_1,H_2$ and $H_3$. Here $B$ denotes the conventional baryon. The mass ordering is $B < H_1 < H_2 < H_3$. In the absense of spin dependent forces all ground states corresponding to a given adiabatic surface are degenerate. The quantum number notation is $(N,\Delta)^{2S+1}J^P$, where $N,\Delta$ is the flavour structure of the wave function (i.e. those of the conventional baryons $N,\Delta$ respectively) and $P$ the parity. } \label{tabqu} \begin{tabular}{|c||r|l|l|} \hline (Hybrid) Baryon & Chirality & $L$ & $(N,\Delta)^{2S+1}J^P$ \\ \hline $B $ & 1 & 0 & $N^2 {\frac{1}{2}}^+, \; \Delta^4 {\frac{3}{2}}^+$\\ $H_1^S $ & 1 & 1 & $N^2 {\frac{1}{2}}^+, \; N^2 {\frac{3}{2}}^+, \; \Delta^4 {\frac{1}{2}}^+, \; \Delta^4 {\frac{3}{2}}^+, \; \Delta^4 {\frac{5}{2}}^+$\\ $H_1^A $ & 1 & 1 & $N^2 {\frac{1}{2}}^+, \; N^2 {\frac{3}{2}}^+$\\ $H_2^S $ & 1 & 1 & $N^2 {\frac{1}{2}}^+, \; N^2 {\frac{3}{2}}^+, \; \Delta^4 {\frac{1}{2}}^+, \; \Delta^4 {\frac{3}{2}}^+, \; \Delta^4 {\frac{5}{2}}^+$\\ $H_2^A $ & 1 & 1 & $N^2 {\frac{1}{2}}^+, \; N^2 {\frac{3}{2}}^+$\\ $H_3^S $ &-1 & 0 & $N^2 {\frac{1}{2}}^-, \; \Delta^4 {\frac{3}{2}}^-$\\ $H_3^A $ &-1 & 0 & $N^2 {\frac{1}{2}}^-$\\ \hline \end{tabular} \end{center} \end{table} The junction can move in three directions, and correspondingly be excited in three ways, giving the hybrid baryons $H_1, H_2$ and $ H_3$. For each junction excitation, it is found that the junction wave function can be realized to be either totally symmetric, or totally antisymmetric under quark label exchange, indicated by $H^S$ and $H^A$ respectively. The quantum numbers of the lowest--lying states that can be constructed on the adiabatic surfaces corresponding to each of the six hybrid baryons are indicated in Table \ref{tabqu}. Since quarks are fermions, the wave function should be totally antisymmetric under quark label exchange, called the Pauli principle. The colour structure of hybrid baryons are taken to be identical to those of conventional baryons, i.e. it is totally antisymmetric under label exchange. This imposes constraints on the combination of flavour and non--relativistic spin $S$ of the three quarks that is allowed. The combinations are indicated in Table \ref{tabqu}. ``Chirality'' gives the behaviour of the junction wave function under reflection in the plane spanned by the quarks, when the positions of the quarks are fixed. Let $L$ be the orbital angular momentum of the quarks and the junction. It is possible to argue that $L=1$ for the ground state $H_1$ and $H_2$ hybrid baryons, while $L=0$ for the ground state conventional and $H_3$ hybrid baryons. The total angular momentum ${\bf J} = {\bf L} + {\bf S}$. Since $L=0$ for ground state conventional and $H_3$ hybrid baryons, $J=S$. Since $L=1$ for ground state $H_1, H_2$ hybrid baryons, $J=\frac{1}{2},\frac{3}{2}$ for $S=\frac{1}{2}$, and $J=\frac{1}{2},\frac{3}{2},\frac{5}{2}$ for $S=\frac{3}{2}$. These assignments are indicated in Table \ref{tabqu}. \section{The potential in which the quarks move} \begin{figure}[t] \vspace{-1.6cm} \begin{centering} \epsfig{file=lb2.ps,width=15cm,angle=0} \vspace{-2.5cm} \caption[x]{Difference of the hybrid and conventional baryon potentials (in GeV) for $\rho=2.881$ GeV$^{-1}$ as a function of $\lambda$ (in GeV$^{-1}$) and $\cos\theta_{\rho\lambda}$.} \label{fig:TheLoop} \end{centering} \end{figure} \vspace{0.3cm} We shall now calculate the junction energy (``adiabatic potential'') for the ground and first excited states of the junction as a function of the quark positions. Define \begin{equation}\label{rho} {\brho} = \frac{{\bf r}_1 - {\bf r}_2}{\sqrt{2}}\hspace{1cm} {\blambda} = \frac{{\bf r}_1+{\bf r}_2-2 {\bf r}_3}{\sqrt{6}}\hspace{1cm} \cos \theta_{\rho\lambda} = \frac{\brho\cdot\blambda}{\rho\lambda} \end{equation} where ${\bf r}_i$ denotes the positions of the quarks. The energy is a function of $\rho$, $\lambda$ and $\theta_{\rho\lambda}$. The procedure for evaluating the conventional baryon potential is as follows. We numerically evaluate $V_B (l_1,l_2,l_3)$ by solving the Schr\"{o}dinger Equation for the junction hamiltonian, $(\frac{1}{2}M_{\mbox{\small eff}}^{\infty}{\bf \dot{r}^2}+V_{\mbox{\small J}})\Psi_B({\bf r}) =V_B (l_1,l_2,l_3)\Psi_B({\bf r}) $, variationally using an ansatz ground state simple harmonic oscillator wave function. $M_{\mbox{\small eff}}^{\infty}$ is the effective mass of the junction in the limit where the flux--tubes between the junction and quarks are continuous strings. $V_{\mbox{\small J}}$ is the linear potential between the junction and the three quarks. The hybrid baryon $H_1$ potential $V_{H_1} (l_1,l_2,l_3)$ is solved using $(\frac{1}{2}M_{\mbox{\small eff}}^{\infty}{\bf \dot{r}^2}+V_{\mbox{\small J}})\Psi_{H_1}({\bf r}) =V_{H_1} (l_1,l_2,l_3)\Psi_{H_1}({\bf r})$ with a first excited state simple harmonic oscillator wave function as an ansatz. The difference between the hybrid and conventional baryon potentials is plotted in Figures \ref{fig:TheLoop} - \ref{fig:TheLoop1}. Since the potentials are functions of $\rho,\lambda$ and $\theta_{\rho\lambda}$, one of the variables is held fixed at a typical value for clarity of presentation. The (hybrid) baryon potential can be seen to increase when $\rho\lambda$ is small. Numerically, the ratio of the hybrid to baryon potential is found to be $1.44 - 1.6$ for all $\rho,\lambda$ and $\theta_{\rho\lambda}$. A preliminary estimate of the ground state $H_1$ hybrid baryon mass of $\sim 2$ GeV has also been made by adding the difference between the hybrid and conventional baryon potentials to the phenomenologically successful baryon potential used in ref. \cite{capstick86} \section{Conclusions} The spin and flavour structure of the six hybrid baryons have been specified. Exchange symmetry constrains the spin and flavour of the (hybrid) baryon wave function. The orbital angular momentum of the low--lying hybrid baryon is argued to be unity. The adiabatic potentials have been calculated numerically. The low--lying hybrid baryon mass has been estimated numerically. \begin{figure}[t] \vspace{-1.8cm} \begin{centering} \epsfig{file=rl2.ps,width=15cm,angle=0} \vspace{-2.8cm} \caption[x]{Difference of the hybrid and conventional baryon potentials for $\cos\theta_{\rho\lambda}=0.002$ as a function of $\rho$ and $\lambda$ (in GeV$^{-1}$).} \label{fig:TheLoop1} \end{centering} \end{figure} \vspace{0.3cm} \vskip 1 cm \thebibliography{References} \bibitem{hadron} P.R. Page, {\it Proc. of ``Seventh International Conference on Hadron Spectroscopy'' (HADRON '97)}, 25--30 August 1997, Upton, N.Y., U.S.A., eds. S.-U. Chung, H.J. Willutzki, p. 553, American Institute of Physics. \bibitem{capstick86} S. Capstick, N. Isgur, {\it Phys. Rev.} {\bf D34} (1986) 2809. \end{document}
\section{Introduction} It is known that the investigation of the transition $\gamma N \rightarrow P_{33}(1232)$, using the experimental data on the pion photoproduction on the nucleons, is connected with the problem of separation of the resonance and nonresonance contributions in the multipole amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$, which carry information on this transition. These amplitudes may contain significant nonresonance contributions, the fact which was clear with obtaining the first accurate data \cite{1,2} on the amplitude $E_{1+}^{3/2}$. The energetic behaviour of this amplitude, in fact, is incompatible with the resonance behaviour. The first investigations of this problem \cite{3,4,5} have shown that it is closely related to the problem of fulfilment of the unitarity condition, which for photoproduction multipole amplitudes (let us denote them as $M(W)$) in the $P_{33}(1232)$ resonance region means the fulfilment of the Watson theorem\cite{6}: \begin{equation} M(W)= \exp[i\delta(W)] |M(W)|. \label{1}\end{equation} Here $\delta$ is the phase of the corresponding $\pi N$ scattering amplitude: \begin{equation} h(W)=\sin[\delta(W)]\exp[i\delta(W)]. \label{2}\end{equation} There are different approaches for the extraction of an information on the $\gamma N \rightarrow P_{33}(1232)$ transition from the pion photoproduction data with different forms of the unitarization of the multipole amplitudes. These approaches can be subdivided into the following groups: the phenomenological approaches \cite{3,4,5,7} including the approaches based on the K-matrix formalism \cite{8,9}, the effective Lagrangian (EL) approaches \cite{10,11,12,13,14} with different phenomenological form of the unitarization of the amplitudes, the dynamical models (DM) \cite{15,16,17,18,19,20,21}, and the approaches based on the fixed-t dispersion relations \cite{22,23,24,25}. In Refs. \cite{24,25} it was shown that fixed-t dispersion relations used within the approach of Refs. \cite{26,27} can be usefull for the separation of the resonance and nonresonance contributions in the amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$. In Sec.2 we specify this separation, making correspondence between the contributions in EL approaches and DM and the solutions of the integral equations for $M_{1+}^{3/2},E_{1+}^{3/2}$, which follow from dispersion relations for these amplitudes within the approach of Refs. \cite{26,27}. These solutions are used in Sec.3 as the input for the description of the results of VPI partial-wave analysis \cite{28} for $M_{1+}^{3/2},E_{1+}^{3/2}$, and for separation of the resonance and nonresonance contributions in these amplitudes. \section{Correspondence between contributions in dispersion relation, effective Lagrangian and dynamical approaches} The results of EL approaches and DM can be interpretated on the diagram language, which is most suitable for comparison with the predictions of existing models, because current hadron models and approaches (quark model, bag model, QCD sum rules ...) operate only with verteces and can not predict the whole amplitudes of the processes. In EL approaches and DM the amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$ are described in terms of the diagrams corresponding to the $N$ exchange in the u-channel, $\Delta$ exchange in the $s$- and $u$- channels, and $\pi$ and $\omega$ exchanges in the $t$- channel. The proper phase of the amplitudes is obtained via taking into account final state interaction. This procedure is carried out in EL models phenomenologicaly using the Olsson \cite{3}, Noelle \cite{29} and K-matrix approaches. In DM the unitarization of the amplitudes is made within some method for calculation of the diagrams corresponding to the final state interaction. These calculations are made using different approaches for fulfilment of relativistic and gauge invariance with different methods of cutoff and incorporation of the off-shell effects in the integrals. The amplitudes corresponding to the $N,~\pi$ and $\omega$ exchanges and to the $\Delta$ exchange in the $u$- channel are real. Let us denote their contribution into $M_{1+}^{3/2}$ and $E_{1+}^{3/2}$ as $M^{NR}(W)$. In the quantum mechanics, rescattering effects in these amplitudes lead to the following replacement (see Ref.\cite{30}, Chapter 9): \begin{equation} M^{NR}\rightarrow M^{NR}_{rescat}=M^{NR}+ \frac{1}{\pi}\frac{1}{D(W)} \int\limits_{W_{thr}}^{\infty} \frac{D(W')h(W')M^{NR}(W')}{W'-W-i\varepsilon}dW'= \label{3}\end{equation} \begin{equation} =\exp [i \delta (W)]\left[M^{NR}(W)\cos \delta (W)+ e^{a(W)}r(W)\right], \label{4}\end{equation} where \begin{equation} r(W)= \frac{P}{\pi} \int\limits_{W_{thr}}^{\infty} \frac{e^{-a(W')}\sin \delta (W')M^{NR}(W')}{W'-W}dW', \label{5}\end{equation} \begin{equation} a(W)= \frac{P}{\pi} \int\limits_{W_{thr}}^{\infty} \frac{W \delta (W')}{W'(W'-W)}dW'. \label{6}\end{equation} In Eq.(\ref{3}) it is supposed that the unitarity condition (\ref{1}) can be used in the whole range of integration, $D(W)$ is the Jost function: \begin{equation} 1/D(W)= \exp\left[\frac{W}{\pi} \int\limits_{W_{thr}}^{\infty} \frac{\delta (W')}{W'(W'-W-i\varepsilon)}dW'\right]= \exp [i \delta (W)]e^{a(W)}. \label{7}\end{equation} The contributions analogous to both terms in Eqs. (\ref{3}),(\ref{4}) exist in all DM. These models reproduce exactly the first term in Eq. (\ref{4}), second one being model dependent and different in different models. In EL approaches the unitarization made via the Noelle and K-matrix ansatzes corresponds to taking into account only the first term in Eq. (\ref{4}) (see Ref. \cite{13}). The unitarization via the Olsson ansatz in these approaches has no analogy with the above formulas. It is interesting that just the first term in Eq. (\ref{4}) determines the nonresonance behaviour of the multipole amplitude $E_{1+}^{3/2}$ (see below the curve 5 in Fig. 2). In the absence of background contribution into $\delta_{1+}^{3/2}$, incorporation of the $\pi N$ rescattering in the resonance parts ($M^R$) of the amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$ leads in the vicinity of $\Delta$ to the following replacements: \begin{equation} M^{R}= \frac{f_{\pi N,\Delta}^0 f_{\Delta,\gamma N}^0} {s -m_{0\Delta}^2} \rightarrow \frac{f_{\pi N,\Delta} f_{\Delta,\gamma N} } {s -m_{\Delta}^2-im_{\Delta}\Gamma_{\Delta}} \equiv \frac{f_{\pi N,\Delta} f_{\Delta,\gamma N}} {m_{\Delta}\Gamma_{\Delta}} sin\delta_R e^{i\delta_R}. \label{8}\end{equation} Here $f_{\pi N,\Delta},~ f_{\Delta,\gamma N},~\Gamma_{\Delta}$ and $m_{\Delta}$ are dressed verteces and $\Delta$ width and mass; the corresponding values containing "0" are bare ones. The modification of $M^R$, due to the presence of the background contribution in $\delta_{1+}^{3/2}$, can be taken into account only phenomenologicaly. One can estimate the magnitude of this modification using the results of Ref.\cite{13} obtained within the Noelle and K-matrix forms of the unitarization of the amplitudes. At the resonance position, where $\delta_{1+}^{3/2}=90^{\circ}$ (for the phase shift analysis of Refs. \cite{31,32} it is $W_R=1.229~GeV$), the unitarization of $M^R$ within these methods leads to the same results; namely, in Eq. (\ref{8}) the replacement $e^{i\delta_R}\rightarrow e^{i\delta_{1+}^{3/2}}$ should be made, $sin\delta_R$ being equal to 1 in the K- matrix approach and to 0.97 in the Noelle approach. This difference in $3\% $ we will consider as the uncertainty of $M^R$ coming from the incorporation of the background contribution into $\delta_{1+}^{3/2}$ at $W=W_R$. Let us turn now to the dispersion relations. Dispersion relations for multipole amplitudes follow from dispersion relations for invariant amplitudes, defined in accordance with the hadron current, which obeys the requirements of the relativistic and gauge invariance and the crossing invariance under the replacement $s \leftrightarrow u$. Let us write these dispersion relations for the multipole amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$ in the form: \begin{equation} M(W)=M^B(W)+M^{high}(W)+ \frac{1}{\pi}\int\limits_{W_{thr}}^{W_{max}} \frac{h^*(W') M(W')}{W'-W-i\varepsilon}dW'+ \frac{1}{\pi}\int\limits_{W_{thr}}^{W_{max}} K(W,W')h^*(W') M(W')dW', \label{9}\end{equation} where we have divided dispersion integrals into two parts: from threshold up to $W_{max}=1.55~GeV$ (the region which is dominated by the $\Delta$ contribution), and from $W_{max}$ up to $\infty$. Such division of the dispersion integrals, with the consideration of $M^{high}(W)$ as a nonsingular function, is possible only in the case, if $h(W)\rightarrow 0$ when $W\rightarrow W_{max}$. This condition was not taken into account in Ref. \cite{23}. By this reason the solutions of integral equations, obtained in \cite{23}, are divergent at $W\rightarrow W_{max}$. For the multipole amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$ one can introduce the condition: $h(W)\rightarrow 0$ when $W\rightarrow W_{max}$, at $W_{max}\simeq1.55~GeV$, because at $W=1.5~GeV$ we have $\delta_{1+}^{3/2}=164^{\circ}$\cite{31,32}. From the $\pi N$ phase shift analyses (see, for example, \cite{31,32}) it is known that the amplitude $h_{1+}^{3/2}$ is elastic in the first integration region; by this reason in the integrals over the region $(W_{thr},W_{max})$ the imaginary parts of the multipole amplitudes are written in the form: $Im~M(W)=h^*(W)M(W)$, which follow from Eqs.(\ref{1},\ref{2}). Therefore, the dispersion relations (\ref{9}) for the amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$ can be considered as integral equations for these amplitudes in the region $(W_{thr},W_{max})$. In Eq. (\ref{9}), $M^B(W)$ is the Born term which corresponds to the $N$ and $\pi$ exchanges, with pseudoscalar coupling for the $NN\pi$ vertex. The corresponding term in EL approaches and DM is obtained using pseudovector coupling for this vertex; it differs from the Born contribution by the nonsingular term which contributes only into the $B_1^{(+,0)}$ Ball amplitude: \begin{equation} B_1^{(+,0}(s,t)=\frac{ge}{4m_N^2}g^{(v,s)}, \label{10}\end{equation} where $m_N$ is the nucleon mass, and \begin{equation} e^2/4\pi=1/137,~g^2/4\pi=14.5,~g^{(v)}=3.7,~g^{(s)}=-0.12. \label{11}\end{equation} The contribution of this term into our final results is negligibly small. $K(W,W')$ is a nonsingular kernel arising from the $u$- channel contribution into the dispersion integral and the nonsingular part of the $s$- channel contribution. In the integrand of the relation (\ref{9}), we did not write the couplings of $M(W)$ to other multipoles; by our estimations their contributions into our final results are negligibly small. The values of the high energy integrals in Eq. (\ref{9}) can be evaluated using the results of analyses of pion photoproduction on nucleons at high energies. In our estimations we have used the results obtained in Ref. \cite{33}, where different variants of the description of these data are considered within the approach based on the Regge poles and cuts. Our estimations have shown that the high energy integrals in Eq. (\ref{9}) can be roughly approximated by the $\omega$ exchange, which contributes to the following Ball amplitudes: \begin{equation} B_6^{(+)}=\frac{2g_{\gamma\omega\pi}g_{\omega NN}}{t-m_{\omega}^2}, ~B_1^{(+)}=m_NB_6^{(+)}, \label{12}\end{equation} where $m_{\omega}$ is the $\omega$ mass, and $g_{\gamma\omega\pi}$ is related to the $\omega\rightarrow\pi\gamma$ decay width by: \begin{equation} \Gamma(\omega\rightarrow\pi\gamma)=\frac {g_{\gamma\omega\pi}^2k^3}{12\pi}, \label{13}\end{equation} $k$ is the pion 3-momentum in the $\omega$ rest frame. From the data on $\Gamma(\omega\rightarrow\pi\gamma)$ \cite{34} we get $g_{\gamma\omega\pi}=0.73~GeV^{-1}$. In Eq. (\ref{12}) we have presented only the contribution corresponding to the vector coupling in the vertex $\omega NN$, because the role of the tensor $\omega NN$ coupling in our final results is negligibly small. For the vector coupling constant we have: $g_{\omega NN}=8-14$ \cite{35}. The results presented below in Figs. 1,2 correspond to the mean value of $g_{\omega NN}$ in this interval. At $K(W,W')=0$, the integral equation (\ref{9}) has a solution in an analitical form (see Refs.\cite{26,27} and the refferences therein): \begin{equation} M_{K=0}(W)=M_{part,K=0}^{B,\omega}(W)+c_MM_{K=0}^{hom}(W). \label{14}\end{equation} Here $M_{part,K=0}^{B,\omega}(W)$ is the particular solution of Eq. (\ref{9}) generated by $M^B$ and $M^{\omega}$. It is described by Eq. (\ref{3}) with the replacement $M^{NR}\rightarrow M^B+M^{\omega}$. With this, in all integrals of Eqs.(\ref{3})-(\ref{7}) at $W'>W_{max}$, one should take $\delta(W')=\pi$. So, $M_{part,K=0}^{B,\omega}(W)$ reproduces the nonresonance contributions into the amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$, generated by the $N,~\pi$ and $\omega$ exchanges, when the final state interaction, caused by the $\pi N$ rescattering in the $\Delta$ region, is taken into account in accordance with Eq. (\ref{3}). $M_{K=0}^{hom}(W)=1/D(W)$ is the solution of the homogeneous equation, which follow from (\ref{9}) at $M^B=M^{\omega}=0$. It enters Eq. (\ref{14}) with an arbitrary weight, i.e. multiplied by an arbitrary constant $c_M$. If, following EL approach and DM, we describe the amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$ in terms of the contributions corresponding to the $N,\Delta,\pi$ and $\omega$ exchanges, then $c_MM^{hom}$ should be considered as the $\Delta$ contribution. In order to obtain the contribution, corresponing to the $\Delta$ exchange in the $s$-channel, one should subtract from $c_MM^{hom}$ the contribution of the $\Delta$ exchange in the $u$-channel. Using final results for the contributions of $c_MM^{hom}$ into $M_{1+}^{3/2},E_{1+}^{3/2}$, one can estimate this contribution. It appeared that the $\Delta$ contribution, corresponding to the $u$-channel, is negligibly small in comparison with $c_MM^{hom}$ and $M_{part}^{B,\omega}(W)$. By this reason, the $\Delta$ contribution in the $s$-channel we identify with $c_MM^{hom}$. Let us note, that our final results correspond to the solutions of the integral equations (\ref{9}) with $K(W,W')\neq 0$, i.e. they satisfy the requirement of the crossing invariance. These solutions were obtained numerically, using the formulas for the amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$ presented in details in Ref. \cite{24}. At $W_{thr}<W<1.5~GeV$, the phase $\delta_{1+}^{3/2}$ was taken in the analitical form \begin{equation} \sin^2\delta ^{3/2}_{1+}=\frac {(4.27q^3)^2} {(4.27q^3)^2+(q_r^2-q^2)^2 [1+40q^2 (q^2-q_r^2)+21.4q^2]^2}, \label{100}\end{equation} which describe well the experimental data from \cite{31,32} with $q_r=0.225~GeV$; $q$ is the 3-momentum of the pion in the $GeV$ units in the $\pi N$ c.m.s. \section{Results and discussion} In this Section we present our results on the description of the data for the multipole amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$ which are extracted with high accuracy from existing experimental data in the partial- wave analysis of Ref. \cite{28}. In the dispersion relation approach, presented in the previous Section, these data should be described as sums of the particular and homogeneous solutions of the integral equations (\ref{9}) for the amplitudes $M_{1+}^{3/2}$ and $E_{1+}^{3/2}$. The particular solutions have definite magnitudes fixed by $M^B$ and $M^{\omega}$, i.e. by the $N,\pi$ and $\omega$ contributions into $M_{1+}^{3/2},E_{1+}^{3/2}$. The solutions of the homogeneous parts of the integral equations (\ref{9}) with $M^B=M^{\omega}$, have definite shapes, fixed by the integral equations, and arbitrary weights. These weights are the only unknown parameters which should be found from the requirement of best description of the data on $M_{1+}^{3/2},E_{1+}^{3/2}$. For this aim we have used fitting procedure. The obtained results together with the data from Ref.\cite{28} are presented in Figs.1,2. In order to demonstrate the role of different contributions, they are presented in these figures separately. The curves 4 and 6 are the particular solutions of Eg.(\ref{9}) generated by $M^B$ and $M^{\omega}$, respectively. They represent the nonresonance contributions into $M_{1+}^{3/2},E_{1+}^{3/2}$, caused by the $N,\pi$ and $\omega$ exchanges. The curves 5 represent the first term in Eq.(\ref{4}) with $M^{NR}=M^B$. They are given in order to demonstrate the difference between the nonresonance contributions, generated by the Born term in the EL aproach of Ref.\cite{13} and our approach. This difference is caused by the second term in (\ref{4}); with this term, the nonresonance contributions, generated by the Born term, satisfy dispersion relations. The curves 3 represent the contributions of the homogeneous solutions, obtained by fitting the weights of these solutions, when the nonresonance contributions are generated by the Born term and $\omega$ exchange. These curves represent the $\Delta$ contributions into $M_{1+}^{3/2},E_{1+}^{3/2}$. As it was mentioned in Sec.2, our estimations have shown that the $u$-channel $\Delta$ contributions are negligibly small in comparison with $s$ -channel ones. By this reason we identify the contributions of the homogeneous solutions (curves 3) with the $\Delta$ exchange in the $s$-channel. The summary results are presented by the curves 1, which correspond to the case, when the nonresonance contributions are caused by the $N,\pi$ and $\omega$ exchanges. It is seen that the agreement with the VPI data is good for both amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$. In order to demonstrate the role of high energy contributions into dispersion integrals which are approximated in our approach by the $\omega$ exchange, we present also the curves 2. They are obtained by fitting the homogeneous solutions, when the nonresonance contributions are generated by the Born terms only. It is seen that the $\omega$ contribution is small; however, in the case of $M_{1+}^{3/2}$ its role in obtaining the good agreement with experiment is important. In Table 1 we present the helicity amplitudes $A^p_{3/2}$ and $A^p_{1/2}$ and the ratio $E2/M1$ for the transition $\gamma N \rightarrow P_{33}(1232)$, which are obtained from the resonance contributions into $M_{1+}^{3/2},E_{1+}^{3/2}$ (the curves 3 in Figs.1,2) at the resonance position. First errors are obtained assuming that the data in Figs.1,2, corresponding to the energy-dependent analysis of Ref.\cite{28} have $2\%$ errors. Second errors come from the uncertainties of the model. They are connected with the cuttof in the dispersion integrals (\ref{9}); with the uncertainties in the $\omega$ contribution; with neglecting the couplings of the multipole amplitudes with each other in (\ref{9}); and with the uncertainties in the extraction of the resonance amplitudes from the curves 3, discussed in the previous Section. Table 1. Helicity amplitudes and the ratio $E2/M1$ for the $\gamma N \rightarrow P_{33}(1232)$ transition\\ \begin{tabular}{|c|c|c|c|} \hline &$A_{1/2}^p(10^{-3}Gev^{-1/2})$&$A_{3/2}^p(10^{-3}Gev^{-1/2})$&$E2/M1(\%)$\\ \hline Resonance contributions,&$-110\pm 2\pm 6$&$-209\pm 4\pm 12$&$-2.2\pm 0.1\pm 0.3$\\ our results&&&\\ \hline Total amplitudes,&$-135\pm 5$&$-250\pm 8$&$-1.5\pm 0.5$\\ Ref.\cite{28}&&&\\ \hline Nonrelativistic&-101&-175&0\\ quark model&&&\\ \hline Relativistic&-111&-207&-2.1\\ quark model \cite{36,37}&&&\\ \hline \end{tabular} \newline In Table 1 we present also the results obtained from the total amplitudes $M_{1+}^{3/2},E_{1+}^{3/2}$ at the resonance position in Ref. \cite{28}. The amplitudes, extracted in such way, are larger than quark model predictions. As is seen from our results, this disagreement is removed due to taking into account the nonresonance background contributions generated by the $N,\pi$ and $\omega$ exchanges. \newline \begin{center} {\large {\bf {Acknowledgments}}} \end{center} I am grateful to I.I.Strakovsky for communications and providing the results of the VPI partial-wave analysis in the numerical form. I also acknowledge communications with B.L.Ioffe and O.Hanstein. \vspace{1cm}
\section{Introduction} In general relativity, with compact rather than asymptotically flat boundary conditions, physical observations are made inside the system that they describe. In quantum theory, the observable quantities are meaningful outside the system that they refer to. It is very likely that quantum gravity must be a ``quantum mechanical relativistic theory". That is, a theory where the observables can be given as self-adjoint operators on a Hilbert space but which are meaningful inside the system that they describe. A rough description of what such a theory involves is the following. Observations made inside the system are closely related to causality in the sense that an inside observer necessarily splits the history of the system into the part that is in the future, the part that is in the past, and --- assuming finite speed of propagation of information --- elsewhere. We may call such observables ``internal observables'', characterised by the requirement that they refer to information that an observer at a point, or a connected region of spacetime, may be able to gain about their causal past. In a previous paper \cite{fm3}, we found that these observables can be described by functors from the partially ordered set of events in the spacetime to the category of sets. Such a functor codes the relationship between the causal structure and the information available to an observer inside the spacetime. This has non-trivial consequences and, in particular, the observables algebra is modified even at the classical level. Internal observables satisfy a Heyting algebra, which is a weak version of the Boolean algebra of ordinary observables. This is still a distributional algebra. Namely, for propositions $P, Q$ and $ R$, if $P\vee Q$ denotes ``$P$ or $Q$'', and $P\wedge Q$ means ``$P$ and $Q$'', then $P\vee(Q\wedge R)= (P\vee Q)\wedge(P\vee R)$. On the other hand, quantum mechanics is linear, and as a result of the superposition principle, quantum mechanical propositions are not distributive. If $P,Q$ and $R$ are projection operators, $P\vee(Q\wedge R)$ is not equal to $(P\vee Q)\wedge(P\vee R)$. Having both quantum mechanics and internal observables in the same theory requires finding propositions that have a non-distributive quantum mechanical aspect and a distributive causal aspect. The aim of the present paper is to define the histories in which such observables may be encountered. We will, therefore, define quantum causal histories, which are histories that are both quantum mechanical and causal. Assuming that a discrete causal ordering (a causal set) is a sufficient description of the fundamental past/future ordering needed to qualify observations as being inside the system, we find that a quantum causal history can be constructed by attaching finite-dimensional Hilbert spaces to the events of the causal set. It is then natural to consider tensor products of the Hilbert spaces on events that are spacelike separated. We define quantum histories with local unitary evolution maps between such sets of spacelike separated events. The conditions of reflexivity, antisymmetry and transitivity that hold for the causal set have analogs in the quantum history which are conditions on the evolution operators. We find that transitivity is a strong physical condition on the evolution operators and, if imposed, implies that the quantum causal histories are invariant under directed coarse graining. If the the causal set represents the universe, quantum causal histories constitute a quantum cosmological theory. Its main notable feature is that there is a Hilbert space for each event but not one for the entire universe. Hence, no wavefunction of the universe arises. A consistent intepretation of quantum causal histories and observatons inside the quantum universe can be provided and will appear in a forthcoming paper \cite{fmls4}. In more detail, the outline of this paper is the following. In section 2 we review causal set histories and provide a list of definitions of structures that can be found in a causal history and which are used in the quantum causal histories. In particular, we concentrate on acausal sets, sets of causally unrelated events. In section 3, we introduce the poset of acausal sets, equipped with the appropriate ordering relation. The definition of quantum causal histories is based on this poset and is given in section 4. The properties of the resulting histories are discussed in section 5. The ordering of a causal set is reflexive, antisymmetric and transitive, conditions which are also imposed on the quantum histories. The consequences of these properties are analyzed in section 6. In particular, we find that transitivity leads to directed coarse-graining invariance. Two classes of quantum causal histories are given as examples in section 7. Up to this point, the causal histories require a choice of a causal set. In section 8, we remove this restriction and provide a sum-over-histories version of quantum causal evolution. The quantum causal histories presented here are consistent, but not all physically meaningful questions can be asked. There are several possibilities for generalisations, some of which we outline in the Conclusions. \section{Causal set histories} A (discrete) causal history is a causal set of events that carry extra structure. For example, the causal histories that were examined in \cite{fmls1, fm2, fmls2} had as events vector spaces spanned by $SU_q(2)$ intertwiners. In two dimensions, an exact model of such a causal history has been proposed by Ambjorn and Loll \cite{AL} and its continuum limit properties have been investigated in \cite{AL,ANRL}. The dynamics of a 3-dimensional causal spin network history model is addressed by Borissov and Gupta in \cite{BG}\footnote{ For pure causal set theories (with no extra structure on the events), we may note recent work by Rideout and Sorkin who derived a family of stochastic sequential growth dynamics for a causal set, with very interesting consequences about the classical limit of pure causal sets \cite{RS}. Further, the dynamics of a toy model causal set using a suitable quantum measure was proposed by Criscuolo and Waelbroek in \cite{CW}.}. In this section, we review the definition of a causal set and provide several derivative definitions which will be used in the rest of this paper. A causal set ${\cal C}$ is a partially ordered set whose elements are interpeted as the events in a history (see \cite{blms,sorkin,dm}). We denote the events by $p,q,r,\ldots$. If, say, $p$ precedes $q$, we write $p\leq q$. The equal option is used when $p$ coincides with $q$. We write $p R q$ when either $p\leq q$ or $p\geq q$ holds. The causal relation is reflexive, i.e.\ $p\leq p$ for any event $p$. It is also transitive, i.e.\ if $p\leq q$ and $q\leq r$, then $p\leq r$. To ensure that ${\cal C}$ has no closed timelike loops, we make the causal relation antisymmetric, that is, if $p\leq q$ and $q\leq p$, then $p=q$. Finally, we limit ourselves to histories with a finite number of events. Given a causal set, there are several secondary structures that we can construct from it and which come in useful in this paper. We therefore list them here (see Figures 1 and 2): \begin{itemize} \item The {\it causal past} of some event $p$ is the set of all events $r\in{\cal C}$ with $r\leq p$. We denote the causal past of $p$ by $P(p)$. \item The {\it causal future} of $p$ is the set of all events $q\in{\cal C}$ with $p\leq q$. We denote it by $F(p)$. \item An {\it acausal set}, denoted $a,b,c,\ldots$, is a set of events in ${\cal C}$ that are all causally unrelated to each other. \item The acausal set $a$ is a {\it complete past} of the event $p$ when every event in the causal past $P(p)$ of $p$ is related to some event in $a$. It is not possible to add an event from $(P(p)-a)$ to $a$ and produce a new acausal set. \item Similarly, an acausal set $b$ is a {\it complete future} of $p$ when every event in the causal future $F(p)$ of $p$ is related to some event in $b$. \item A {\it maximal antichain} in the causal set ${\cal C}$ is an acausal set $A$ such that every event in $({\cal C}-A)$ is causally related to some event in $A$. \end{itemize} Similar definitions as the above of past, future, complete past, and complete future for a single event can be given for acausal sets: \begin{itemize} \item The causal past of $a$ is $P(a)=\bigcup_i P(q_i)$ for all $q_i\in a$. Similarly, the causal future of $a$ is $F(a)=\bigcup_i F(q_i)$ for all $q_i\in a$. \item An acausal set $a$ is a complete past of the acausal set $b$ if every event in $P(b)$ is related to some event in $a$. \item An acausal set $c$ is a complete future of $b$ if every event in $F(b)$ is related to some event in $c$. \end{itemize} Furthermore, \begin{itemize} \item Two acausal sets $a$ and $b$ are a {\it complete pair} when $a$ is a complete past of $b$ and $b$ is a complete future of $a$. \item Two acausal sets $a$ and $b$ are a {\it full pair} when they are a complete pair and every event in $a$ is related to every event in $b$. \item Two acausal sets $a$ and $b$ {\it cross} when some of the events in $a$ are in the future of $b$ and some are in its past. \end{itemize} \begin{figure} \centerline{\mbox{\epsfig{file=PastFuture.eps}}} \caption{This is a very small causal history. $P(p)$ is the causal past of the event $p$ while $F(p)$ is its causal future. The acausal set $a$ is a complete past for $p$, and $b$ is a complete future. } \label{PastFuture} \end{figure} \begin{figure} \centerline{\mbox{\epsfig{file=completepair.eps}}} \caption{The acausal sets $a$ and $b$ are a complete pair.} \label{completepair} \end{figure} \section{The poset of acausal sets} The set of acausal sets within a given causal set ${\cal C}$ is a partially ordered set if we define the relation $a\preceq b$ to mean that $a$ is a complete past of $b$ and $b$ is a complete future of $a$. Reflecting the properties of the underlying causal set, the relation $\preceq$ is reflexive, transitive and antisymmetric. Let us call this poset ${\bf A}$. It is on this poset of acausal sets that we will base the quantum version of the causal histories. Its properties, therefore, are important constraints on the corresponding quantum history. The main property of ${\bf A}$ that characterises the kind of quantum theory we will obtain in this paper is that, given acausal sets $a, b$ and $c$, the following holds ($R$ means either $\preceq$ or $\succeq$): \begin{equation} \mbox{If }aRc, bRc,\mbox{ and }a, b\mbox{ do not cross, then } aRb. \end{equation} That is, given some acausal set $c\in{\bf A}$, all the acausal sets that are related to $c$ are also related to each other --- except if they happen to cross. If $a$ and $b$ are ``too close'' to each other they may cross and then cannot be related by $\preceq$. This means that for the chosen $c$ there is not a unique complete pair sequence. In selecting one of the possible sequences we need to make repeated choices of which of two crossing acausal sets to keep. \section{Quantum causal histories} \label{qch} We will now construct the quantum version, $Q{\bf A}$, of the poset ${\bf A}$. We will regard an event $q$ in the causal set as a Planck-scale quantum ``event'' with a Hilbert space $H(q)$ that stores its possible states. We require that $H(q)$ is finite-dimensional, which is consistent with the requirement that our causal sets be finite. Choose an acausal set $a=\{q_1,q_2,\ldots,q_n\}$ in ${\bf A}$. Since all $q_i\in a$ are causally unrelated to each other, standard quantum mechanics dictates that the Hilbert space of $a$ is \begin{equation} H(a)=\bigotimes_{i=1}^n H(q_i). \end{equation} That is, we have a tensor product Hilbert space in $Q{\bf A}$ for each acausal set in ${\bf A}$. When two acausal sets are related, $a\preceq b$, there needs to be an evolution operator between the corresponding Hilbert spaces: \begin{equation} E_{ab}:H(a)\longrightarrow H(b). \label{eq:E} \end{equation} We will impose one more condition on the causal histories that will make their present treatment simpler. We will only consider posets ${\bf A}$ with the following property: \begin{equation} \mbox{ If }a\preceq b,\qquad\mbox{dim}\ H(a)=\mbox{dim}\ H(b). \end{equation} This restriction is particularly convenient since it allows us to simply regard $H(a)$ and $H(b)$ as isomorphic and require that $E_{ab}$ is a unitary evolution operator. The poset of acausal sets ${\bf A}$ is reflexive, transitive and antisymmetric. We would like to maintain these properties of the causal ordering in the quantum theory as analogous conditions on the evolution operators. (In other words, we want the quantum causal history to be a functor from the poset ${\bf A}$ to the category of Hilbert spaces.) The analogue of reflexivity is the existence of an operator $E_{aa}={\bf 1}_a:H(a)\rightarrow H(a)$ for every acausal set $a$. $E_{aa}$ has to be the identity because any other operator from $H(a)$ to itself would have to be a new event. Transitivity in ${\bf A}$ implies that \begin{equation} E_{bc} E_{ab}=E_{ac} \label{eq:transitivity} \end{equation} in $Q{\bf A}$. We will return to transitivity and its implications for ${\bf A}$ and $Q{\bf A}$ in section 6. At each event $q$, there is an algebra of observables, the operators on $H(q)$. An observable $\widehat{O}_a$ at $a$ becomes an observable $\widehat{O}_b$ at $b$ by \begin{equation} \widehat{O}_b={E_{ab}}\widehat{O}_a E_{ab}^\dagger. \end{equation} This completes the definition of the causal quantum histories we are concerned with. In the next section we discuss the evolution of states which is allowed in such histories. Then, in section 6, we will come back to the imposition of transitivity on the evolution operators, a strong condition that dictates the form of the resulting histories and their quantum cosmology interpretation. \section{Quantum evolution in $Q{\bf A}$} In this section we discuss the consequences of the definitions of quantum histories given above. \subsection{Products of complete pair sequences} Evolution maps between complete pairs that are themselves causally unrelated to each other may be composed in the standard way, like tensors. That is, consider complete pairs $a\preceq b$ and $c\preceq d$, with $a$ and $b$ unrelated to $c$ or $d$. Then construct the acausal sets $a\cup c$ and $b\cup d$, which form a new complete pair: $(a\cup c)\preceq (b\cup d)$. The evolution operator on the composites, \begin{equation} E_{(a\cup c)(b\cup d)}:H(a)\otimes H(c) \longrightarrow H(b)\otimes H(d), \end{equation} is the product of the operators on the two pairs, \begin{equation} E_{(a\cup c)(b\cup d)}=E_{ab}\otimes E_{cd}. \end{equation} \subsection{Projection operators} The causal structure of ${\bf A}$ means that a projection operator on $H(a)$ propagates to the future of $a$ in the following way. A projection operator \begin{equation} P_a:H(a)\rightarrow V(a) \end{equation} that reduces $H(a)$ to a subspace $V(a)$, can be extended to a larger acausal set $a\cup c$. On $H(a)\otimes H(c)$, it is the new projection operator \begin{equation} P_{a\cup c}=P_a\otimes{\bf 1}_c :H(a)\otimes H(c)\longrightarrow V(a)\otimes H(c). \end{equation} By using the evolution operator $E_{(a\cup c)(b\cup d)}=E_{ab}\otimes E_{cd}$ on the enlarged projection operator we obtain a projection operator $P_{b}\otimes 1_{d}$ on the future acausal set $b\cup d$. \subsection{Evolution in ${\bf A}$ can be independent of ${{\cal C}}$} \label{rho_problem} Consider a complete pair $a\preceq b$ in which $a=a_1\cup a_2$ and $b=b_1\cup b_2$. The corresponding Hilbert spaces are \begin{equation} H(a)=H(a_1)\otimes H(a_2)\qquad\mbox{and} \qquad H(b)=H(b_1)\otimes H(b_2), \end{equation} and $E_{ab}$ is the evolution operator that corresponds to the causal relation $a\preceq b$. Choose some state $|\psi\rangle\in H(a_1)$ by acting on $H(a_1)$ with the projection operator $|\psi\rangle\langle\psi|$. This implies that, in $H(a)$, we have chosen the state $|\psi\rangle\otimes|\psi_{a_2}\rangle$, for some $|\psi_{a_2}\rangle\in H(a_2)$, using the projection operator $(|\psi\rangle\langle\psi|)\otimes{\bf 1}_{a_2}$. We can use $E_{ab}$ on this state to obtain the state \begin{equation} |\psi_b\rangle=E_{ab}\left(|\psi\rangle\otimes|\psi_{a_2}\rangle \right) \end{equation} in $H(b)$. If, for any reason, we need to restrict our attention to $b_2$, we can trace over $b_1$ to find that the original state $|\psi\rangle\in H(a_2)$ gives rise to the density matrix: \bea \rho_\psi(b_2)&=&\mbox{Tr}_{b_1}|\psi_b\rangle\langle\psi_b\|\\ &=&\mbox{Tr}_{b_1}\left[E_{ab}\left(|\psi\rangle \langle\psi|\otimes{\bf 1}_{a_2}\right)E_{ab}^\dagger\right]. \label{eq:rho} \eea At this point, the following question arises. What if $a_1$ is not in the causal past of $b_2$ and, for example, we have these causal relations: \begin{equation} \begin{array}{c}\mbox{\epsfig{file=violation.eps}}\end{array} \nonumber \end{equation} Does the evolution we defined on $Q{\bf A}$ violate causality in the underlying causal set ${\cal C}$? This question illuminates several features of the quantum causal histories. The very first thing to note is that we get the same acausal poset ${\bf A}$ for many different causal sets. The operator $E_{ab}$ refers to ${\bf A}$ and does not distiguish the different possible underlying causal sets. There is a simple solution to the above apparent embarassment. Instead of promoting the events of the causal set to Hilbert spaces, we may attach the Hilbert spaces to the edges, and the evolution operators to the events. An event in the causal set, then, becomes an evolution operator from the tensor product of the Hilbert spaces on the edges ingoing to that event, to the tensor product of the outgoing ones. Since the set of ingoing and the set of outgoing edges to the same event are a full pair (i.e.\ a complete pair in which all events in the past acausal set are related to all the events in the future one), the above problem will not arise. Conceptually, this solution agrees with the intuition that events in the causal set represent change, and, therefore, in the quantum case they should be represented as operators. In section \ref{ex2}, we discuss the example of quantum causal histories with the Hilbert spaces on the edges for trivalent causal sets. \subsection{Propagation by a density matrix requires a complete pair} \label{aw} According to (\ref{eq:rho}), given a state $|\psi\rangle\in H(a)$, we can obtain the density matrix $\rho_\psi(b_{2})$ for the acausal set $b_{2}$ in the future of $a$. This uses the fact that $b_{2}$ is a subset of an acausal set $b$ that forms a complete pair with $a$. Consider this configuration: \begin{equation} \begin{array}{c}\mbox{\epsfig{file=aw.eps}}\end{array} \nonumber \label{eq:awset} \end{equation} The initial acausal set is $a=\{p_1,p_2\}$. The acausal set $w=\{p_3, p_4\}$ is in the future of $a$. Given $|\psi\rangle\in H(a)$, can we obtain a density matrix on $H(w)$? The answer is no, since there is no acausal set that contains $w$ and is maximal in the future of $a$. (The problem is the $p_2\leq p_7$ relation.) There are, therefore, acausal sets in the future of $a$ which cannot be reached by the evolution map $E$. \section{Directed coarse-graining} The main idea in this paper is that the quantum version of some causal history is a collection of Hilbert spaces connected by evolution operators that respect the structure of the poset ${\bf A}$ we started with. For this reason, in section 4, we imposed reflexivity, transitivity and antisymmetry to the operators $E_{ab}$. Transitivity is a strong condition on the quantum history. One should keep in mind that it was first imposed on causal sets because it holds for the causal structure of Lorentzian spacetimes. Significantly, it does not just encode properties of the ordering of events but also the fact that a Lorentzian manifold is a {\it point set}. In general relativity, an event is a point and this has been imported in the causal set approach. To analyse this a little further, let us introduce a notation that indicates when two events $p$ and $q$ are related by a shortest causal relation, i.e., no other event occurs after $p$ and before $q$. This is the {\it covering relation}: \begin{quote} $\bullet$ The event $q$ {\it covers} $p$ if $p\le q$ and there is no other event $r$ with $p\le r\le q$. We denote this by $p\to q$. \end{quote} The following are worth noting. For a finite causal set, transitivity means that the order relation determines, and is determined by, the covering relation, since $p\le q$ is equivalent to a finite sequence of covering relations $p=p_1\to p_2 \ldots\to p_n=q$. On the other hand, in the continuum (for example the real line ${\bf R}$) there are no pairs $p,q$ such that $p\to q$ \cite{DavPri}. Hence, in a continuum spacetime, it is simply not meaningful to consider an ordering that is not transitive. Non-transitive ordering requires distinguishing between the covering relations and the resulting transitive ones. This distinction is not possible in the continuum case. In short, for events that are points, it is sensible to expect that if $p$ leads to $q$ and $q$ to $r$, then $r$ is in the future of $p$. If, however, the events were (for example) spacetime regions of some finite volume, with overlaps, then it is unclear whether transitivity would hold. (See also section 2.4 in \cite{CJI2}.) In the causal histories we consider here, an event is a Hilbert space. It is, therefore, an open question whether it is sensible to impose reflexivity, transitivity and antisymmetry on the ordering of the Hilbert spaces. We choose to first impose them, then find what the implications are, and if they are unphysical, go back and check which of the three conditions should not be maintained in a quantum causal ordering. On the positive side, there is a very interesting advantage to maintaining transitivity. Using (\ref{eq:transitivity}), we have the benefit of a {\it directed coarse-graining} invariance of the quantum history. For example, if we are handed \begin{equation} \begin{array}{c}\mbox{\epsfig{file=cg1.eps}}\end{array} \nonumber \end{equation} and need to go from $H(p_1)\otimes H(p_2)$ to $H(p_3)\otimes H(p_4)\otimes H(p_5)$, we can reduce it to \begin{equation} \begin{array}{c}\mbox{\epsfig{file=cg2.eps}}\end{array}. \nonumber \end{equation} Clearly, several initial graphs will give the same coarse-grained graph. We return to this in section \ref{sums}. It is very interesting to note that the coarse-graining implied by transitivity can be used to improve the propagation of density matrices that we discussed in section \ref{aw}. We can coarse-grain the causal set depicted in (\ref{eq:awset}) by considering the events $p_3,p_4,p_5,p_6,p_7$ to be the acausal set $\bar{w}$. This is a coarse-graining in the sense that we ignore any causal relations between these events. We then obtain: \begin{equation} \begin{array}{c}\mbox{\epsfig{file=cg3.eps}}\end{array}. \nonumber \end{equation} In this causal set, $a$ and $\bar{w}$ are a complete pair. It is then possible to use an evolution operator $E_{a\bar{w}}$ to take the state $|\psi\rangle\in H(a)$ to a state in $H(\bar{w})$ and then trace over $(\bar{w}-w)$ to obtain a density matrix on $H(w)$. \section{Examples} In this section, we provide two examples of the quantum causal histories we defined in section \ref{qch}. \subsection{Discrete Newtonian evolution} A discrete Newtonian history is a universe with a preferred time and foliation. It is represented by a poset ${\bf A}$ which is a single complete pair sequence: \begin{equation} \ldots a_n\longrightarrow a_{n+1}\longrightarrow a_{n+2}\ldots . \end{equation} The corresponding quantum history is then \begin{equation} \ldots H_n\longrightarrow H_{n+1}\longrightarrow H_{n+2} \ldots. \end{equation} with all the Hilbert spaces isomorphic to each other. We can denote by $E$ the evolution operator from $H_n$ to $H_{n+1}$. Evolution from $H_n$ to $H_m$ is then given by \begin{equation} E_{nm}=E^{m-n}. \end{equation} We may compare this universe to the standard one in quantum theory. There, there is a single Hilbert space for the entire universe and evolution is given by the unitary operator $e^{iH\delta t}$. In the above, there is a sequence of identical and finite-dimensional Hilbert spaces. Since evolution is by discrete steps, we may set $\delta t=1$. Then $E=e^{iH}$, for some hermitian operator $H$ and $E_{nm}=e^{i(m-n)H}$. \subsection{A planar trivalent graph with Hilbert spaces on the edges} \label{ex2} This example is a history with multifingered time \cite{fm2}. It is a planar trivalent graph with finite-dimensional Hilbert spaces living on its {\it edges}. Trivalent means either two ingoing and one outgoing edges at a node, or two outgoing and one ingoing. We exclude nodes with no ingoing or no outgoing edges. From a given planar trivalent causal set ${\cal C}$, we can obtain what we will call its {\it edge-set}, ${\cal E}{\cal C}$. This is a new graph which has the covering relations of ${\cal C}$ (the edges, not including any transitive ones) as its nodes. The covering relations in ${\cal C}$ are also ordered and these relations are the edges in ${\cal E}{\cal C}$. Figure \ref{CandEC} shows an example of a causal set and its edge-set. We now take the poset ${\bf A}$ of ${\cal E}{\cal C}$ and construct a quantum history from it by assigning Hilbert spaces to the nodes of ${\cal E}{\cal C}$. The very interesting property of ${\cal E}{\cal C}$ is that it can be decomposed into pieces, generating evolution moves that take two events to one, or split one event into two. That is, ${\cal E}{\cal C}$ can be decomposed to these two full pairs: \begin{equation} \begin{array}{c}\mbox{\epsfig{file=updown.eps}}\end{array}. \nonumber \end{equation} (This is generally the case, not only for trivalent graphs.) Such a decomposition is not possible in some general ${\cal C}$ and therefore, in view of the problem we encountered in \ref{rho_problem} above, ${\cal E}{\cal C}$ provides an advantage over ${\cal C}$. To be able to employ unitary evolution operators, we need $\mbox{dim}\ H(e_1)=\mbox{dim}\ H(e_2)+\mbox{dim}\ H(e_3)$ and $\mbox{dim}\ H(e_4)+\mbox{dim}\ H(e_5)=\mbox{dim}\ H(e_6)$. One can check that this can be done consistently for all the events in the causal set. Although ${\cal C}$ is trivalent, the nodes of ${\cal E}{\cal C}$ have valence 2,3, or 4. The list of possible edges in ${\cal C}$ and the corresponding nodes in ${\cal E}{\cal C}$ is: \begin{equation} \begin{array}{c}\mbox{\epsfig{file=valences.eps}}\end{array} \nonumber \end{equation} There is a substantial interpretational difference between Hilbert spaces on the (covering) causal relations and on the events. A state space on the edge $p\leq q$ is most naturally interpeted as ``the state space of $p$ as seen by $q$''. If there are two edges coming out of $p$, to $q$ and $r$, then there are two Hilbert spaces in ${\cal E}{\cal C}$, interpeted as the Hilbert space of $p$ as seen by $q$ and the Hilbert space of $p$ as seen by $r$. On the other hand, a Hilbert space placed on the event $p$ is absolute, in the sense that is independent of who is observing it. \begin{figure} \centerline{\mbox{\epsfig{file=CandEC.eps}}} \caption{A causal set and its edge-set.} \label{CandEC} \end{figure} \section{Summing over histories} \label{sums} The quantum causal histories we discussed in this paper require a fixed poset ${\bf A}$. This is also necessary in the treatment of observers in a classical causal set universe in \cite{fm3}. Since there is no physical reason that some causal set should be preferred over all others, this restriction is unappealing. In this section we outline a ``sum-over-histories'' version of the evolution in section \ref{qch}, which also applies to \cite{fm3} and can be used in any further work on quantum observers inside the universe. An acausal set is a set of points. We have considered causal sets where all events have at least one ingoing and one outgoing edge. Then, that $a,b$ are a complete pair $a\preceq b$ means that there is a directed graph with the set $a$ as its domain and the set $b$ as its codomain. Let us denote this graph by $\gamma(a,b)$. A graph with $b$ as its codomain and one with $b$ as its domain may be composed. When ${\bf A}$ is given, there is one known graph $\gamma(a,b)$ connecting $a$ and $b$. If ${\bf A}$ is not fixed, we can sum over all graphs that we can fit between $a$ and $b$ (that have a finite number of nodes). This leads to a sum-over-histories version of the evolution in section \ref{qch}. Let us call $E_{ab}^\gamma$ the evolution operator (as in equation (\ref{eq:E})) when the underlying graph is $\gamma(a,b)$. The transition amplitude from a state $|\psi_a\rangle\in H(a)$ to a state $|\psi_b\rangle\in H(b)$ for this particular graph is \begin{equation} A_\gamma=\langle\psi_b|E^\gamma_{ab}|\psi_a\rangle. \end{equation} If the graph is not fixed, we may sum over all the possible ones: \begin{equation} A_{ab}=\sum_\gamma \langle\psi_b|E^\gamma_{ab}|\psi_a\rangle. \label{eq:A} \end{equation} Now note that transitivity defines equivalence classes of graphs between given acausal sets. Here is an example of a series of graphs that, by transitivity, have the same causal relations as far as $a$ and $b$ are concerned. Specifically, $p_1,p_2,p_3\leq p_4$ and $p_3\leq p_4,p_5$ in all of these graphs: \begin{equation} \begin{array}{c}\mbox{\epsfig{file=directed.eps}}\end{array} \nonumber \end{equation} All of the above correspond to the same evolution operator in (\ref{eq:A}). It is intriguing to compare this to the triangulation invariance of topological quantum field theory. Transitivity can be intepreted as a {\it directed} triangulation invariance. We will return to this in future work. \section{Conclusions} We saw that it is possible to promote a causal set to a quantum one by taking the events to be finite-dimensional Hilbert spaces. It is natural to consider tensor products of the Hilbert spaces on events that are spacelike separated. This led us to replace the causal set by the poset of acausal sets ${\bf A}$. We defined quantum histories with local unitary evolution maps between complete pairs in ${\bf A}$.\footnote{Discrete evolution by unitary operators is also considered by P.\ Zizzi, in work to appear.} We explored several of the features of these causal histories. A property of ${\bf A}$ is that it splits into distinct sequences of complete pairs. Consequently, this places restrictions on which Hilbert spaces can be reached from a given one. The conditions of reflexivity, antisymmetry and transitivity that hold for ${\bf A}$ were imposed on the quantum history as conditions on the evolution operators. The most interesting consequences were from transitivity, which gave rise to invariance under directed coarse-graining. This needs further investigation, since the physical assumption behind transitivity is pointlike events. It will be interesting to consider events which are extended objects and find what ordering is suitable in this case. We were able to tensor together the Hilbert spaces for two acausal sets when they are spacelike separated and have no events in common. However, it is natural to consider cases where an acausal set is a subset of a larger one. It appears that, if we need to use acausal sets, we should enrich ${\bf A}$ with the inclusion relation, that is, use a poset with {\it two} ordering relations, causal ordering and spacelike inclusion. While this is straightforward in the plain causal set case, it becomes tricky in the quantum histories. For example, we may start from the acausal sets $a=\{q_1,q_2\}$ and $a^\prime=\{q_1,q_2,q_3\}$. In the poset ${\bf A}$ we have $a\subset a^\prime$. In $Q{\bf A}$, the corresponding state spaces are $H(a)=H(q_1)\otimes H(q_2)$ and $H(a^\prime)=H(q_1)\otimes H(q_2)\otimes H(q_3)$. However, there is no natural way in which $H(a)$ is a subspace of $H(a^\prime)$. That is, the set inclusion relation is not directly preserved in the quantum theory. In the quantum histories we discussed, we restricted the past and future Hilbert spaces to have the same dimension. This is the simplest case and needs to be generalised. A related issue is the properties of the individual Hilbert spaces. For example, if we employ the causal spin network models of \cite{fmls2}, the evolution operators should respect the $SU(2)$ invariance of the state spaces. More generally, a discrete quantum field theory toy model can be constructed by inserting matter field algebras on the events. This will be addressed in future work. Finally, we set up $Q{\bf A}$ having in mind a functor from a poset to Hilbert spaces, taking the elements and arrows of the poset into Hilbert spaces and operators which preserve the properties of the original poset, i.e.\ reflexivity, antisymmetry and transitivity. It is also possible to use graphs between sets of events (rather than a fixed causal set), as outlined in section \ref{sums}. In this case, the quantum causal histories become similar to topological quantum field theory except, importantly, they are directed graphs (or triangulations). The coarse-graining invariance relations can be calculated for given fixed valence of the covering relations. On the interpretational side, the main thing to note is that the causal history is a collection of Hilbert spaces which itself is not a Hilbert space. According to quantum theory, we can take tensor products of events that are not causally related. We cannot, for example, take the tensor product of all the Hilbert spaces in the history to be the Hilbert space of the entire history.\footnote{Taking tensor products of all Hilbert spaces in the history is, however, exactly what is done in the consistent histories approach of Isham \cite{CJI}.} As a result, the causal quantum cosmology is not described in terms of a wavefunction of the universe. Individual events (or observers on the events) can have states and wavefunctions but the entire universe does not. Further discussion of the interpretation of the quantum causal histories will appear in \cite{fmls4}. \section*{Acknowledgment} I am grateful to Chris Isham for his comments on transitivity and the pointlike structure of spacetime. This paper has benefited from discussions with Lee Smolin on quantum theory and the interpetation of such histories as quantum cosmologies. I am also grateful to Sameer Gupta and Eli Hawkins for discussions on causal histories. This work was supported by NSF grants PHY/9514240 and PHY/9423950 to the Pennsylvania State University and a gift from the Jesse Phillips Foundation. \newpage
\section{Introduction}\label{intro} \indent\indent This talk was initially meant to be given by Tae-Sun Park since he is the one doing most of the work -- with a little help from three of us -- but the organizers asked me (MR) to present the paper instead. In giving this talk, I would like to distinguish between the statements I am making for which the other members of the collaboration should not be held responsible and those that are endorsed by all of us. The former will be addressed (mostly in footnotes) by ``I" and the latter by ``We." Since the early attempt to apply effective field theories to nuclear physics~\cite{weinberg,pmr,vankolck}, there have been many papers written on the subject, the most recent development of which is being summarized in this meeting~\cite{thismeeting}. In confronting Nature with effective field theories in nuclei -- one of the main themes of this meeting, one has had to be content mostly with {\it postdictions}, genuine {\it predictions} being harder to come by. The reason is simply that effective field theories involve, at each order of power counting, a certain number of parameters in the effective Lagrangian. It is believed that those parameters are in principle calculable from first principles for a given scale ({\it e.g.}, lattice QCD~\cite{lepage}) but in practice, they have to be fixed by experimental data. Once the parameters are so fixed, the Lagrangian can then be used to make predictions for other processes that involve the same parameters. Up to date, however, most of the calculations involved fixing of the parameters and only rarely could one predict and check the prediction by experiments. In this talk, we would like to present a genuine prediction based on the formulation of an effective field theory that we have developed during the last few years. Since our formulation is available in the literature, we will not dwell on the details of the formalism but present the essence of the arguments while stressing the possible caveats involved. We illustrate how well postdictions can be made and then how to make predictions for two specific processes, one for which data are available and hence the theory can be tested immediately and the other for which data are not yet available but will be forthcoming, offering a marvelous possibility for an honest prediction totally unbiased by available experiments. \section{The Chiral Filter}\label{chiralfilter} \indent\indent An early attempt predating the advent of QCD and effective field theories in nuclear physics was made in 1978 under a conjecture called the ``chiral filter hypothesis"\cite{KDR}. Based on current algebras, it was argued that corrections to the single-particle transitions in nuclei for the isovector M1 operator and the weak axial charge operator should be dominated by one-soft-pion-exchange two-body currents. As a corollary, if the soft-pion exchange two-body currents are suppressed either by symmetry or kinematics as for instance in the isoscalar EM current or the Gamow-Teller operator ({\it i.e.}, the space component of the axial current), the chiral filter says that corrections to the single-particle operator cannot be given by a few controlled terms, thereby making the calculations highly model-dependent. This hypothesis that appeared to be somewhat ad hoc at the time it was proposed turned out to be justified in the context of chiral perturbation theory~\cite{pmr}: The soft-pion exchange term is indeed the leading order correction in the chiral counting wth the next-to-leading correction calculable but strongly suppressed. There are two consequences of this chiral filter that survive in the context of modern effective field theories. One is a nontrivial postdiction and the other is an interesting prediction that has been largely confirmed. \subsection{A strategy for effective field theory (EFT)}\label{strategy} \indent\indent As it stands, there are effectively two ``alternative" ways of power counting in setting up effective field theories for two-nucleon systems. One is the original Weinberg scheme~\cite{weinberg} in which the leading four-Fermi contact interaction and a pion-exchange are treated on the same footing in calculating the ``irreducible" graphs for a potential that is to be iterated to all orders in the ``reducible" channel. The power counting is done only for the irreducible vertex. The other is the ``power divergence subtraction" (PDS) scheme~\cite{KSW} in which only the leading (nonderivative) four-Fermi contact interaction is iterated to all orders with the higher-order contact interactions and the pion exchange treated perturbatively. While the PDS scheme is perhaps more systematic in the power counting, we believe that the Weinberg scheme is not only consistent with the strategy of EFT but also, in the sense developed below, more flexible and predictive with possible errors committed due to potential inconsistency in the power counting generically suppressed. In our work, this scheme is adopted. Iterating to all orders in the reducible channel with the irreducible vertex is equivalent to solving the Schr\"odinger equation with a corresponding potential. This then suggests that we use, in calculating response functions to slowly varying electroweak fields that we are interested in, those wave functions computed with so-called ``realistic potentials," the prime example being the Argonne $v_{18}$ potential~\cite{argonne} (called in short Av18). In fact this procedure of mapping effective field theory to realistic wave functions -- a hybrid approach~\cite{KK} -- for two-nucleon response functions employed previously by us~\cite{pmr} has recently been justified by means of a cutoff regularization~\cite{PKMR}. van Kolck has presented a similar argument in support of such a hybrid procedure~\cite{vk}. Now in this framework, the power counting reduces simply to a chiral counting in the irreducible vertex for the current as the current appears only once in the graphs. We are then allowed to separate the current matrix element of interest into the single-particle and exchange-current terms, with the single-particle matrix element given entirely by that with the Av18 wave function and the exchange current contribution -- given by the matrix element with the same wave function -- from operators computed in standard baryon chiral perturbation theory. The old soft-pion exchange term figuring in the chiral filter -- if not suppressed -- is just the leading order contribution in this series. \subsection{Unpolarized $np$ capture} \indent\indent A good example of postdictions that follow from the above scheme and that will be a basis for a genuine prediction described below is the process \begin{eqnarray} n+p\rightarrow d+\gamma\label{np} \end{eqnarray} where both nucleons are unpolarized and the incoming neutron has a thermal energy~\footnote{The relative momentum in the center of mass is $\sim 3.4\times 10^{-3}\ \ {\rm MeV}$.}. This process has been computed to the next-to-next-to-leading order (NNLO) in the chiral counting in the scheme described above for the current~\cite{pmr2}. The theoretical cross section $\sigma_{th}=334\pm 3\ \ {\rm mb}$ which agrees with the experiment within the error bar consists of the leading contribution, $305.6\ \ {\rm mb}$, coming from the single-particle matrix element given by the Av18 wave function and the remainder from the exchange current dominated by the soft one-pion exchange according to the chiral filter. Two points are worth noting in this result. First, this is a bona-fide calculation and {\it not} a fit: Within the scheme adopted here, there are no free parameters. Secondly as shown in \cite{PKMR}, the single-particle matrix element has a negligible uncertainty, so that the error in the theory is entirely attributed to the uncertainty in the exchange-current operator associated with the short-distance part of the interactions that cannot be accessed by chiral perturbation theory. This part introduces a scale and renormalization-scheme dependence and only those results that are not sensitive to this short-distance uncertainty can be trusted. In the framework in which the realistic wave functions figure, the short-distance scale is set by the cutoff proportional to $r_C^{-1}$ where $r_C$ is the ``hard-core radius" that removes the part of the wave function for $r\roughly< r_c\neq 0$. The net effect of this cutoff is that in addition to cutting the radial integrals in the coordinate space, it removes {\it all} zero-range terms in the current operator including zero-ranged counter terms. The $1 \%$ error bar assigned to the theory for (\ref{np}) represents the uncertainty in this cutoff procedure. This procedure can be justified for the process in question by using a cutoff $\sim r_C^{-1}$ and showing that the counter term ``killed" by the hard core is in that uncertainty range. Given that the four-Fermi counter term is removed by the hard core which may be viewed as exploiting a scheme dependence, there are no more parameters in the theory~\footnote{See appendix B in the second reference of \cite{pmr2} for the zero-ranged counter term in question. In the framework of \cite{KSW} which avoids scheme dependence at the expense of predictiveness, this counter term is a non-negligible parameter, rendering a bona-fide calculation infeasible~\cite{npKSW}.}. This procedure, familiar to nuclear physicists, will be referred to as ``hard-core cutoff scheme" (or HCCS in short). Below we will see that when the chiral filter does not apply, the removal of the zero-range counter terms by the simple hard-core cutoff may be suspect. \subsection{Prediction for the axial-charge transitions in nuclei}\label{axial} \indent\indent The EFT scheme described in section \ref{strategy} makes a rather clean prediction which has not yet been adequately appreciated in nuclear physics community. The chiral filter idea~\cite{KDR}, by now validated in chiral perturbation theory~\cite{pmr}, predicts that in the nonrelativistic regime, the axial charge transition matrix element for the $\beta$ decay process in nuclei \begin{eqnarray} A (J^{\pm})\rightarrow A^\prime (J^\mp) +e +\nu; \ \ \ \Delta T=1\label{axialcharge} \end{eqnarray} should receive a huge one-soft-pion exchange current correction, amounting to more than 50 \% of the single-particle matrix element. The prediction for this is quite robust and firm with very little nuclear model dependence: possible corrections from higher chiral order terms -- which are calculable a priori -- are estimated to be less than 10 \% of the leading soft-pion terms~\cite{pmr,pkt}. How do we go about checking this prediction against experiments? It is not possible to test this in the two-nucleon systems for which the effective field theory is most extensively developed: the matrix element for (\ref{axialcharge}) cannot be measured in few-nucleon systems. The measurement can only be made in heavier nuclei. This means that to compare with experiments, we would have to account for the effect of density or multi-nucleon interactions on the transition. This introduces a subtlety which is interesting in its own right. We will show below that there is rather compelling evidence that both the chiral filter and density effect are confirmed~\footnote{The issue of effective field theory for hadrons immersed in a hadronic medium with baryon density $\rho$ was one I (MR) would have liked to discuss in this meeting but it is out of the scope of this presentation. For a recent discussion on this subject, we simply refer to the article~\cite{frs}. For the present problem, the net effect is the presence of what is known as ``Brown-Rho (BR) scaling" $\Phi (\rho)$~\cite{BR}.}. Experimentalists extract the axial charge matrix element ${\cal M}$ from their experiments and then write~\cite{warburton} \begin{eqnarray} \epsilon_{MEC}={\cal M}_{measured}/{ M}_{th-1b} \end{eqnarray} where the numerator is the total ``measured" value or more precisely the value extracted from experiments and the denominator is the {\it theoretical} single-particle matrix element of the single-particle current whose constants are {\it unrenormalized} by medium. The model dependence of the denominator -- and to some extent the numerator -- makes this quantity not entirely empirical, thus open to controversy among theorists: It would seem that the ``experimental'' $\epsilon_{MEC}$ would in practice depend upon the model for the wave function used for the sing-particle matrix element. On the other hand, our claim is that the theoretical expression for this quantity is well-defined and free of model dependence~\footnote{ Several people in the audience voiced doubt as to whether the experimental test I discussed is truly valid. One of the reasons given is the model dependence of the so-called empirical information in $\epsilon_{MEC}$. This objection is to some extent valid and needs to be carefully examined. The presently employed procedure is to calculate the single-particle matrix element with the ``best" nuclear wave functions of the single-particle axial charge operator whose coupling constants are {\it unrenormalized} by medium. The question then is what does one mean by ``best" wave function? Here we should stress that in the case in question, there is a reasonably satisfactory answer. Let's take Warburton's analyses~\cite{warburton}. Although his analyses do involve shell-model wavefunctions, the semi-empirical nature of the effective transition operator method he adopted gives his results much more robustness than people normally expect from shell-model calculations. That is, after including the core polarization effects in the form of rescaling the single-particle matrix elements, there is in fact very little room for changing nuclear physics input in his analyses. As a measure of the basic soundness of his approach, we mention the following fact. As far as nuclear dynamics is concerned, the rank-zero and rank-one first-forbidden transitions share the same feature, but the exchange currents for them have very different behavior. The rank-one operators coming from the space component of the axial current has only small exchange currents whereas the time component of the axial current that contributes to the rank-zero operator is expected to have a {\it huge} exchange current. Now, Warburton's calculations reproduce very well the strengths of all the rank-one matrix elements with no further adjustments whereas, for the rank-zero matrix elements, he found it clearly necessary to introduce the ``empirical'' enhancement factor $\epsilon_{MEC}$. The other reason (which is theoretical) is that the BR scaling is so far implemented in the limit of infinite nuclear matter and the question arises regarding the finite size effect of the nuclei measured. My answer to this objection is that whereas experimentally the quantity $\epsilon_{MEC}$ may perhaps be somewhat model-dependent, theoretically, however, it is very well defined: It involves the BR scaling factor $\Phi$ which is smoothly varying for density up to $\rho=\rho_0$, so what matters is the average density involved and the quantity ${\cal R}$, the ratio of the matrix element of the exchange current over that of the single-particle operator, which is quite insensitive to nuclear models provided the same wave functions are used for both.}. Let ${\cal M}_{total}$ denote the total theoretical axial-charge matrix element and ${\cal M}_{n}$ with $n=1, 2$ be the matrix element of the $n$-body operator effective in medium. Then the prediction is that \begin{eqnarray} {\epsilon_{MEC}}^{th}&=&{\cal M}_{total}/M_1=\Phi^{-1} (1+{\cal R}), \label{eth}\\ {\cal M}_{total}&=&{\cal M}_1 +{\cal M}_2 \end{eqnarray} with \begin{eqnarray} {\cal R}={\cal M}_2/{\cal M}_1= \left(M_2/M_1\right) (1+{\cal O} (10^{-1})) \end{eqnarray} where $M_n$ is the matrix element of the $n$-body current with its basic coupling constants ({\it e.g.}, $f_\pi$, $g_A$ etc) {\it unrenormalized} by medium. This relation follows within our HCCS (hard-core cutoff scheme) provided one assumes that all light-quark hadron mass $M$ {\it except for the pion mass} scales in medium as \begin{eqnarray} M^\star/M\approx f_\pi^\star/f_\pi\approx \Phi (\rho). \end{eqnarray} The expression (\ref{eth}) was first derived by Kubodera and Rho~\cite{KR91} and later corrected~\cite{pkt,frs} \footnote{There was an error in the original derivation due to the fact that the non-scaling of the pion mass, indicated by both theory and experiment, was not properly taken into account.}. Now what we need is the value for $\Phi$ and ${\cal R}$ for a range of nuclei that have been measured. The range of nuclear density involved is $1/2 \roughly< \rho/\rho_0 \roughly< 1$ where $\rho_0$ is the nuclear matter density. For this range, we know from information gotten from giant dipole resonances and QCD sum rules in medium~\cite{frs} that $\Phi$ goes as $\sim 1/(1+0.28 (\rho/\rho_0))$, \begin{eqnarray} 0.78\roughly< \Phi\roughly< 0.88 \ \ {\rm for}\ \ 1/2\roughly< \rho/\rho_0\roughly< 1. \end{eqnarray} The two-body current in the ratio ${\cal R}$ is given within 10\% by a soft-pion exchange which is completely fixed by chiral symmetry. The ratio turns out to be extremely {\it insensitive} to nuclear models used to compute the matrix elements and depends only slightly on density. It comes out to be \footnote{Owing to the chiral filter applicable to this process, the uncertainty in our hard-core cutoff scheme, {\it i.e.}, ``killing'' zero-range counter terms, is reflected in the 10\% uncertainty in ${\cal R}$ mentioned above.} \begin{eqnarray} 0.43\roughly< {\cal R}\roughly< 0.61 \ \ {\rm for}\ \ 1/2\roughly< \rho/\rho_0\roughly< 1. \end{eqnarray} This gives the range \begin{eqnarray} 1.63\roughly< \epsilon_{MEC}^{th}\roughly< 2.06 \ \ {\rm for}\ \ 1/2\roughly<\rho/\rho_0 \roughly< 1. \end{eqnarray} A formula as simple as (\ref{eth}) must have a simple and clean test. It must be readily confirmed or infirmed. We claim that there is a strong empirical support for this prediction. Indeed this prediction can be compared with the presently available ``empirical" values~\cite{minamisono,baumann,vangeert,warburton} \begin{eqnarray} \epsilon_{MEC}^{exp}&=& 1.64\pm 0.05\ \ (A=12) \ ; 1.62\pm 0.05\ \ (A=50) \ ;\nonumber\\ && 1.95\pm 0.05\ \ (A=205) \ ; 2.05\pm 0.05\ \ (A=205\sim 212). \end{eqnarray} We further suggest that these constitute evidence for {\it both} (1) the gigantic enhancement predicted by the chiral filter, ranging from 46\% to 61 \% and (2) the additional enhancement predicted by BR scaling, ranging from 20 \% to 45 \%. There is of course the caveat that individual transitions must be subject to some finite-size effects but the point is that what one is probing here is a generic bulk property of nuclear matter -- which to us is the most interesting part of the story. \section{Polarized Neutron-Proton Capture: A New Probe} \indent\indent We will now go outside of those processes protected by the chiral filter and make a genuine prediction even though the process, unprotected by the chiral filter, could be highly suppressed. One would think that such a prediction is out of the scope of effective field theories but surprisingly, it turns out not to be the case in our EFT scheme. \subsection{Selection rules} \indent\indent Consider the process \begin{eqnarray} \vec{n}+\vec{p}\rightarrow d+\gamma\label{polnp} \end{eqnarray} where now both the target proton and the projectile neutron are polarized. This process is being measured at the Institut Laue-Langevin (ILL) in Grenoble by M\"uller {\it et al}.~\cite{mueller} The interest in this experiment is that with the polarized target and the polarized beam, one can measure small matrix elements that are overwhelmed by the dominant isovector M1 matrix element when averaged over polarization. To see this, look at the quantum numbers involved in the process. The initial state of (\ref{polnp}) at very low energy we are dealing with can be in either $^1S_0$ or $^3S_1$ channel. The Fermi-Dirac statistics requires that the former must be in $T=1$ and the latter in $T=0$ where $T$ is the isospin. The final nuclear state in (\ref{polnp}) is the deuteron which is in $^3S_1$ or $^3D_1$ with $T=0$. There are then three relevant transition matrix elements with the emission of a soft photon, {\it i.e.}, isovector M1, isoscalar M1 (which we shall denote from now on as $\overline{\rm M1}$ to distinguish it from the isovector M1) and isoscalar E2. We shall adopt the convention of M\"uller et al~\cite{mueller} and write the transition amplitude as \begin{eqnarray} \langle \psi_d(M_d), \gamma({\hat k} \lambda) | {\cal T}| \psi_{np}(s_p, s_n)\rangle = \chi^\dagger_{1 M_d}\, {\cal M}({\hat k}, \lambda)\, \chi_{s_p} \chi_{s_n} \end{eqnarray} with \begin{eqnarray} {\cal M}({\hat k} \lambda) = \sqrt{4\pi} \frac{\sqrt{v_n}}{2 \sqrt{\omega} A_s} \, \left[ i ({\hat k} \times \epsilon^*)\cdot ({\vec \sigma}_1-{\vec \sigma}_2)\, {\mbox{M1}({}^1S_0)} \right. \nonumber \\ \left. - i ({\hat k} \times \epsilon^*)\cdot ({\vec \sigma}_1+{\vec \sigma}_2)\,\frac{{\mbox{M1}({}^3S_1)}}{\sqrt{2}} + ({\vec \sigma}_1\cdot{\hat k} {\vec \sigma}_2\cdot \epsilon^* + {\vec \sigma}_2\cdot{\hat k} {\vec \sigma}_1\cdot \epsilon^*) \frac{{\mbox{E2}({}^3S_1)}}{\sqrt{2}} \right]\label{amp} \end{eqnarray} where $M_d$ and $\lambda$ are respectively the polarizations of the deuteron and the photon, ${\hat k}$ is the unit momentum vector of the photon, $\omega$ its energy, ${\vec \epsilon} \equiv {\vec \epsilon}(\vec k \lambda)$, $v_n$ is the velocity of the neutron and $A_s$ is the deuteron normalization factor $A_s\simeq 0.8850\ {\rm fm}^{-1/2}$. In the way defined, the quantities ${\mbox{M1}({}^1S_0)}$, ${\mbox{M1}({}^3S_1)}$ and ${\mbox{E2}({}^3S_1)}$ all have a dimension of ${\rm fm}$ and the cross section for the unpolarized $np$ system takes the form \begin{eqnarray} \sigma_{unpol}= \abs{{\mbox{M1}({}^1S_0)}}^2 + \abs{{\mbox{M1}({}^3S_1)}}^2 + \abs{{\mbox{E2}({}^3S_1)}}^2\,.\label{xsection} \end{eqnarray} The first term is the isovector M1 contribution, the second the isoscalar M1 and the last the isoscalar E2. As we shall see below, the second and third terms are strongly suppressed compared to the first, $\sim O(10^{-6})$, so the unpolarized cross section cannot ``see'' these terms. \subsection{Polarization observables} \indent\indent In order to see those small isoscalar terms, one measures polarization observables, {\it i.e.}, the photon circular polarization $P_\gamma$ and the photon anisotropy $\eta$. For an unpolarized proton and a polarized neutron with the polarization vector $\vec{P}_n$, $P_\gamma$ is given by \begin{eqnarray} P_\gamma = \frac{|{\vec P}_n|}{2}\, \frac{2 \sqrt{2} ({\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2} - \rE) + ({\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}+\rE)^2}{1 + {\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}^2 + \rE^2}\label{pgamma} \end{eqnarray} where we have defined the ratios \begin{eqnarray} {\cal R}_{\rm M1} &\equiv& \frac{{\mbox{M1}({}^3S_1)}}{{\mbox{M1}({}^1S_0)}}, \nonumber \\ {\cal R}_{\rm E2} &\equiv& \frac{{\mbox{E2}({}^3S_1)}}{{\mbox{M1}({}^1S_0)}}.\label{ratio} \end{eqnarray} Since these ratios turn out to be $\sim 10^{-3}$, eq.(\ref{pgamma}) simplifies with high accuracy to \begin{eqnarray} P_\gamma \simeq |{\vec P}_n|\, \sqrt{2} ({\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2} - \rE). \end{eqnarray} The anisotropy measures the fully polarized $np$ system involving the polarization of the target and projectile nucleons and is given by \begin{eqnarray} \eta&=& \frac{I(90^\circ) - I(0^\circ)}{ I(90^\circ) + I(0^\circ)} \nonumber \\ &=& pP\, \frac{{\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}^2 + \rE^2 - 6 {\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2} \rE}{ 4 (1 - pP) + (4 + pP) ({\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}^2 + \rE^2) + 2 pP\, {\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2} \rE} \end{eqnarray} where $I$ is the photon intensity, \begin{eqnarray} pP\equiv {\vec P}_p\cdot{\vec P}_n\, \end{eqnarray} and the angle in $I$ measures the photon direction with respect the spin polarization of the neutron and the proton. Note that unless the factor $(1-pP)\sim 0$ the anisotropy $\eta$ will be quadratically suppressed while $P_\gamma$ is linear in the ratio. If however $(1-pP)\sim 0$, then the anisotropy could be substantial, supplying an additional formula that would allow one to extract the two ratios (\ref{ratio}). The purpose of the experiment is to determine these two ratios. The quantity $P_\gamma$ has already been measured before by Bazhenov {\it et al}~\cite{russian}, so the aim of the ILL experiment~\cite{mueller} is to measure the anisotropy $\eta$. \subsection{Doing EFT} \indent\indent As mentioned in section \ref{chiralfilter}, the current involved here is not protected by the chiral filter. This means that the soft-pion exchange which is entirely given by chiral symmetry considerations cannot contribute. The corollary to the chiral filter hypothesis would then suggest that we might be opening Pandora's box. To our surprise, this does not seem to be the case for the problem at hand. \vskip 0.3cm $\bullet$ {\bf Power Counting} \vskip 0.3cm Since the isovector M1 operator is calculated~\cite{pmr} very accurately to ${\cal O} (Q^3)$ relative to the single-particle operator ({\it i.e.}, M1$(^1S_0)=5.78\pm 0.03$ fm in the notation of (\ref{xsection}) which comes at ${{\cal O}} (Q^{-2}$)) \footnote{Unless otherwise specified, we will always give counting relative to the leading single-particle operator.}, we will focus on the isoscalar operators here. \begin{figure}[htbp] \begin{center} \epsfig{file=EFT-INT-fig1.epsi} \caption[gene]{\protect \small Generic diagrams for the two-body isoscalar current ${\cal B}_{\rm 2B}^\mu$. The solid circles include counter-term insertions and (one-particle irreducible) loop-corrections. The wiggly line stands for the external field (current) and the dashed line the pion. One-loop corrections for the pion propagator and the $\pi NN$ vertex are of course to be included at the same order.}\label{fig1} \end{center} \end{figure} We assume that as in the case studied so far~\cite{PKMR}, given the accurate wave functions, the leading single-particle matrix elements are accurately given for both $\overline{\rm M1}$ ({\it i.e.}, isoscalar M1) and E2 operators. The power counting that we adopt~\cite{weinberg,pmr,PKMR} then shows that while the ratio $\langle {\rm 2-body}\rangle/\langle {\rm 1-body}\rangle$ goes like ${{\cal O}} (Q^1)$ for the operator protected by the chiral filter ({\it e.g.}, the isovector M1), the ratio for the isoscalar operators goes like ${{\cal O}} (Q^n)$ with $n\geq 3$, so naively one would expect a suppression by two orders of power counting. Now the question is: How well can we pin down such two-body terms and higher-order corrections to them? \begin{figure}[htbp] \begin{center} \epsfig{file=EFT-INT-fig2.epsi} \caption[deviate]{\protect \small One-loop graphs that contribute to the ${\cal B} \pi NN$ vertex where ${\cal B}$ is the isoscalar current. They give rise to ${\cal O} (Q^4)$ and higher corrections to the leading order (LO) one-body term. \label{fig2} } \end{center} \end{figure} In our scheme, there are two classes of irreducible two-body terms in the current operator. The class A term consists of graphs with the one-pion exchange involving an $NN\pi\gamma$ vertex (see Fig.\ref{fig1}a) and the class B term consists of graphs with two- or more-pion exchanges (see Fig.\ref{fig1}b). Since there is no chiral-filter-protected one-pion exchange in the isoscalar vector current, the class A term receives its leading contribution from one-loop corrections (see Fig.\ref{fig2}). The class B term is generically given by the loop graphs of Fig.\ref{fig3}. The loop and derivatives (in Fig.\ref{fig3}c) account for the additional power suppression. The graphs in Figs.\ref{fig2} and \ref{fig3} typically contribute to the isoscalar M1 operator ($\overline{\rm M1}$) and E2 operator at ${\cal O} (Q^4)$. This is however not the whole story to ${\cal O} (Q^4)$, for there are so-called ``counter terms'' that can come in at ${\cal O} (Q^3)$. There are in fact two such terms in the case that we are concerned with. One such term is a one-pion exchange graph in Fig.\ref{fig1}a with the ${\cal B}\pi NN$ vertex given by a finite counter term. This term that contributes to the $\overline{\rm M1}$ operator but not to the E2 operator is dominated by the $\gamma\rho\pi$ coupling in the anomalous parity component of the effective chiral Lagrangian, {\it i.e.}, the Wess-Zumino term, which is connected with the Adler-Bell-Jackiw triangle anomaly. This term is known and so brings in no unknown parameters. The other is a four-Fermi counter term, call it $g_4$, in Fig.\ref{fig1}b contributing only to $\overline{\rm M1}$. The coefficient of this counter term, not known a priori, needs to be fixed in the usual way. Note that there are no ${\cal O} (Q^3)$ counter term contributions to the E2. \begin{figure}[htbp] \begin{center} \epsfig{file=EFT-INT-fig3.epsi} \caption[deviate]{\protect \small One-loop graphs that contribute to the two-body baryonic currents. They come at ${\cal O} (Q^4)$ and higher order relative to the LO one-body term. All possible insertions of the external line are understood. \label{fig3} } \end{center} \end{figure} \vskip 0.3cm $\bullet$ {\bf There is No Parameter in the Theory} \vskip 0.3cm We shall now argue that while there remains one parameter undetermined by the theory it can be gotten rid of in either of the two ways in which physics at short distance is treated. We first note that there are no unknown parameters in the class A diagrams: As mentioned, the ${\cal O} (Q^3)$ counter term is fixed by the Wess-Zumino term and the loop graphs are completely calculable to ${\cal O} (Q^4)$ without any parameters. In the class B diagrams, there are two parameters, call them $V_i$ with $i=1, 2$, associated with Fig.\ref{fig3}c. One of them, $V_1$, contributes to both $\overline{\rm M1}$ and E2 and the other, $V_2$, only to $\overline{\rm M1}$. We find that the $V_1$ term plays no role in $\overline{\rm M1}$ or E2, as it is suppressed by some power of $Qr_c \ll 1$. The upshot of all this is that we are finally left with the four-Fermi counter term $g_4$ and $V_2$. Furthermore, both of them are associated with zero-range terms in the coordinate space. The combination that appears here will be called, in short, the $g_4``+" V_2$ term. So we can combine them with the ${\cal O} (Q^3)$ zero-ranged operator that comes from the one-pion exchange term with the Wess-Zumino vertex (we shall call this in short the ${\cal O} (Q^3)$ WZ term), thereby reducing them effectively to only one parameter in $\overline{\rm M1}$. The E2 operator is completely free of parameters to ${\cal O} (Q^4)$. Now in the HCCS, the single parameter of the theory does not figure since the contact operator that is multiplied by the parameter gets suppressed by the hard core. But there is a possible caveat here: Since the leading order of this term is ${\cal O} (Q^3)$ whereas the finite loop corrections are of ${\cal O} (Q^4)$, it is not obvious that the HCCS should be reliable in the present case. One might suspect that here short-distance physics could intervene more strongly than when the chiral filter is operative. There are two possible remedies to this problem. One is to implement the operator-product-expansion (OPE) factorization in the wave function suggested by Lepage~\cite{lepagelecture} and the other is to use a cutoff $\sim r_c^{-1}$ in Fourier-transforming the current operators into the coordinate-space form. We shall use the latter which is a lot simpler. We use an equivalent method which is to replace the delta function in all zero-ranged operators by the delta-shell form \begin{eqnarray} \delta (r)\rightarrow \delta (r-r_c).\label{deltashell} \end{eqnarray} This procedure allows the ${\cal O} (Q^3)$ contact operator to contribute in $\overline{\rm M1}$, hence allowing to fix the unknown $g_4``+" V_2$ parameter by fitting the deuteron magnetic moment. It turns out however that this contact term is dominated by the known ${\cal O} (Q^3)$ WZ term, so in practice, the unknown parameter plays only a minor role here. We shall call this scheme the ``modified hard core cutoff scheme." \subsection{Our predictions} \indent\indent We shall now make {\it our} predictions. \vskip 0.3cm $\bullet$ {\bf Hard Core Cutoff Scheme (HCCS)} \vskip 0.3cm With the hard core in the wave function, the parameter-dependent term is killed, so we can now predict the deuteron magnetic moment $\mu_d$, the quadrupole moment $Q_d$, and the ratios ${\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}$ and $\rE$. There must of course be some dependence on the hard core size which enters as a cutoff but the consistency of EFT requires that the cutoff dependence be small. \vskip 0.2cm {\bf Deuteron magnetic moment} $\mu_d$: There is a strong cancellation between the ${\cal O} (Q^3)$ and ${\cal O} (Q^4)$ two-body terms, leaving the one-body term essentially uncorrected: The net two-body correction is found to be less than 0.7\% of the one-body term. The predicted values for $r_c=0.01, 0.2, 0.4, 0.6, 0.8$ fm are (in units of nuclear magneton) \begin{eqnarray} \mu_d = 0.8408,\ 0.8443,\ 0.8426,\ 0.8407,\ 0.8390. \end{eqnarray} The experimental value is $\mu_d^{exp}=0.8574$. \vskip 0.2cm This small discrepancy will be exploited later to fix the one parameter that figures when the zero-range operator is not killed by the hard core as in MHCCS. {\bf Deuteron quadrupole moment} $Q_d$: The two-body correction is equally tiny, less than 0.6\% and more or less independently of $r_c$ within the range $0.01\roughly< r_c\roughly< 0.8$ fm. The result is (in unit of fm$^2$) \begin{eqnarray} Q_d= 0.2710\label{Qd} \end{eqnarray} to be compared with the experiment $Q_d^{exp}=0.2859\ {\rm fm}^2$. It appears that the 5\% discrepancy found here cannot be understood in low-order effective field theories. \vskip 0.2cm ${\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}$: The two-body terms of ${\cal O} (Q^3)$ and ${\cal O} (Q^4)$ in $\overline{\rm M1}$, separately, are of the same magnitude as the one-body term, so although naively suppressed in the power counting, there is no genuine suppression according to the hierarchy of the order. However there is a considerable cancellation between the two higher-order terms, leaving the correction to be between 9\% and 25\% of the single-particle value. The predicted values for $r_c=0.01, 0.2, 0.4, 0.6, 0.8$ fm are \begin{eqnarray} -{\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}\times 10^3=0.869,\ 0.788,\ 0.826,\ 0.871,\ 0.887. \end{eqnarray} $\rE$: As in the quadrupole moment, the two-body correction to the E2 matrix element is small $\roughly< 0.4\%$, so the result is essentially given by the one-body term. Thus within the $r_c$ range considered, the result is, independently of $r_c$, \begin{eqnarray} \rE\times 10^{3}= 0.242.\label{RE} \end{eqnarray} {\bf Photon circular polarization} $P_\gamma$: With the above values for ${\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}$ and $\rE$, the predicted values for $r_c=0.01, 0.2, 0.4, 0.6, 0.8$ fm are (for $|\vec{P}_n|=1$) \begin{eqnarray} -P_\gamma\times 10^3=1.57,\ 1.46,\ 1.51,\ 1.57,\ 1.60. \end{eqnarray} {\bf Photon anisotropy} $\eta$: Here the prediction is extremely sensitive to the value of the polarization $pP$. We will quote for three cases: $pP=1$ (the ideal case), $pP=0.96$ (the highest polarization that may be reached), and $pP= (0.5)^2$ (a case most certainly accessible to the experiment). The predicted values for $r_c=0.01, 0.2, 0.4, 0.6, 0.8$ fm are \begin{eqnarray} \eta [pP=1]&=& 0.57, \ 0.61, \ 0.59,\ 0.57, \ 0.56,\nonumber\\ \eta [pP=0.96]\times 10^{5}&=& 1.3, \ 1.1, \ 1.2, \ 1.3, \ 1.3,\nonumber\\ \eta [pP=(0.5)^2]\times 10^7&=& 1.7, \ 1.5, \ 1.6, \ 1.7, \ 1.8. \end{eqnarray} \vskip 0.3cm $\bullet$ {\bf Modified Hard Core Cutoff Scheme (MHCCS)} \vskip 0.3cm We shall apply the ``smoothing'' (\ref{deltashell}) to the delta function and account for the term carrying information on the single parameter available in the theory. We have the possibility to fix the constant by ``fine-tuning'' it to the deuteron magnetic moment, that is, attributing the 5\% discrepancy in $\mu_d$ from the experimental value to the counter term containing the ${\cal O} (Q^3)$ WZ term and the $g_4``+" V_2$ term. As mentioned, in practice, this term is completely dominated by the former (in fitting the deuteron magnetic moment); therefore the unknown constant plays only a minor role in the resulting isoscalar M1 operator that is to be used to compute the $\overline{\rm M1}$ matrix element. While $\mu_d$ is no longer predicted in this scheme (since it is used to pin down the small $g_4``+" V_2$ term), the deuteron quadrupole mement $Q_d$ remains unmodified from eq.(\ref{Qd}). The ratios ${\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}$ and $\rE$ are of course predicted. \vskip 0.2cm ${\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}$: There turns out be a remarkable $r_c$ independence for this quantity. In fact, in the range $0.01\leq r_c\leq 0.8$ fm, the result is the same: \begin{eqnarray} -{\cal R}_{\rm M1}}\def\rE{{\cal R}_{\rm E2}\times 10^3= 0.500. \end{eqnarray} $\rE$: This quantity remains unchanged from the HCCS, (\ref{RE}), \begin{eqnarray} \rE\times 10^3= 0.242. \end{eqnarray} $P_\gamma$ and $\eta$: Independently of $r_C$ in the range $0.01\leq r_c \le 0.8$ fm, we find \begin{eqnarray} -P_\gamma\times 10^3=1.05 \end{eqnarray} and \begin{eqnarray} \eta [pP=1]&=& 0.80,\nonumber\\ \eta [pP=0.96]&=& 0.62\times 10^{-5},\nonumber\\ \eta [pP=(0.5)^2]&=& 0.86\times 10^{-7}. \end{eqnarray} \section{Conclusion: Call for a Bet} \indent\indent The values we have are not the final ones. First of all, they will have to be rechecked more thoroughly, and secondly, given that the isoscalar matrix elements are so suppressed relative to the isovector matrix element, it may be necessary to take into account the usually negligible isospin violation (both in the interaction and electromagnetic radiative corrections). In any event, we shall give our preliminary predictions here with the warning that they are subject to further changes. We expect to be able to publish a paper with our final numbers in the near future with the above caveats taken into account~\cite{pkmrILL}. In Table \ref{prediction} are summarized our predictions. Recall that there are two schemes for treating the zero-ranged counter terms. One scheme, referred to as hard core cutoff scheme (HCCS), is the usual nuclear physics practice to kill the delta function terms in the operator. The physics so purged from them is presumably shifted to the finite matrix elements affected by the ``correlation hole." In this case there are no more parameters left to account for the small deviation from the experimental data in $\mu_d$ and $Q_d$ and the results would depend on $r_c$; the dependence would of course be weak if the procedure were consistent with the premise of EFT. The other scheme, called modified hard core cutoff scheme (MHCCS), exploits the non-vanishing of zero-range operators to fix one parameter available in the theory by fitting the deuteron magnetic moment, which then determines completely the isoscalar M1 operator to ${\cal O} (Q^4)$. In this scheme, there is practically no $r_c$ dependence and the convergence of the higher-order terms is assured by the procedure (specifically, there is no difference between the ${\cal O} (Q^3)$ calculation and the ${\cal O} (Q^4)$ calculation once $\mu_d$ is fit). \begin{table}[ht] \begin{center} \begin{tabular}{|r||r|r|} \hline Hard-core scheme &HCCS &MHCCS \\ \hline $10^3\times P_\gamma$ & $-1.5\pm 0.1$ & $-1.1$ \\ \hline $\eta[pP=1]$ & $0.58\pm 0.03$ & 0.80 \\ $10^5\times \eta [pP=0.96]$ & $1.2\pm 0.1$ & 0.62 \\ $10^7 \times \eta[pP=0.25]$ & $1.6\pm 0.1$ & 0.86 \\ \hline \hline $10^3 \times {\cal R}_{M1}$ & $-0.84\pm 0.05$ & $-0.51$ \\ $10^3 \times {\rE}$ & 0.24& 0.24\\ \hline \end{tabular} \end{center} \caption{\protect \small Predictions using two schemes for implementing the hard core: HCCS (hard-core cutoff scheme) in which zero-ranged operators are ``killed'' by the short-range correlation function in the wave function and MHCCS (modified hard core cutoff scheme) in which the delta function of zero-range operators is ``smoothed'' to the delta shell form. The ``error bar'' represents the variation over the range $0.01\ {\rm fm}\leq r_c\leq 0.8\ {\rm fm}$. No error bar means that there is no $r_c$ dependence. }\label{prediction} \end{table} There is only one polarization data available at the moment, namely, the photon circular polarization measured by the Russian group~\cite{russian} \begin{eqnarray} P_\gamma^{exp}=(-1.5\pm 0.3)\times 10^{-3}.\label{russian} \end{eqnarray} As it stands, this experiment seems to favor the HCCS result. However the prediction in which we have more confidence is that of MHCCS, which puts it somewhat lower than the experimental value (\ref{russian}). Clearly additional measurements will be needed to confirm (\ref{russian}) or improve on it. \vskip 0.3cm $\bullet$ {\bf The Bet} \vskip 0.3cm {\it We would now like to invite other workers in the field -- particularly those who are in this audience, who have worked out consistent and systematic power countings -- to make similar predictions and participate in the bet for the best prediction, that is, the one that agrees best with the experimental result that is forthcoming. As for our prediction, the most interesting possibility would be that the experiment simply disagrees with our two versions of the hard-core scheme. That would sharpen our chiral filter conjecture and bring in totally new physics.} \footnote{ At the time of my talk, I made a specific proposal for the prize for the winner and I confirm that proposal here. I would propose that each participant should be willing to offer to the winner a bottle of one of the {\it best} wine of his/her country. As for me, I would offer a bottle of superb French wine, Chateau Mouton-Rothschild, a ``premier grand cru class\'e.''} \section*{Acknowledgments} \indent\indent We would like to thank the organizers of this meeting for the invitation to participate in the exciting debate and Thomas M\"uller for discussions and correspondence on the on-going experiment at the ILL. \newpage \section*{References}
\section{Introduction} \label{sec:intro} Simulations of the thermal evolution of neutron stars confronted with the soft X-ray and extreme UV observations of the photon flux emitted from their surface provide one of the most useful diagnostics of the physical processes operating in the dense interior of such stars. The early evolution of a neutron star is completely dominated by the cooling via neutrino emission; at this stage the star effectively radiates away the initial content of its thermal energy. In the later epoch one faces a competition between cooling and heating processes, where the heating may even dominate in the late stages of the thermal evolution. Eventually, they reach a thermal equilibrium phase, where the heat generated in the star is radiated away at the same rate from the stellar surface. The relative importance of the various elementary processes contributing to cooling is fairly well understood (e.g. \cite{Richardson82}; \cite{VanRiper91}; \cite{Umeda94}; \cite{Schaab95a}; \cite{Page97a}). The investigations performed so far have however the deficit that they did not sufficiently clarify the relative weight of the different heating processes. The majority of these investigations concentrated on few preferred dissipation processes (e.g \cite{Shibazaki89}; \cite{Cheng92a}; \cite{Sedrakian93}; \cite{Reisenegger95a}) with the exception of the work of Van Riper (1991)\nocite{VanRiper91}, who treats the heating in general phenomenological terms. A meaningful comparison of the individual processes studied in these papers is strongly hampered by the fact that these investigations were performed for different microphysical inputs and different levels of sophistication concerning the cooling simulation code, preventing one from drawing definitive conclusions about the relative importance of the individual heating processes. The aim of this paper is a comparative analysis of the impact of different heating processes on the thermal evolution of neutron stars. Our simulations take into account the most recent developments concerning the microphysical input and employ both relativistic and non-relativistic equations of state of superdense matter. We perform the cooling calculations for a fully relativistic, evolutionary simulation code. Although internal heating is generally azimuthally asymmetric, we shall use an one-dimensional cooling code for almost all simulations. To demonstrate the reliability of this approximation, a two-dimensional calculation is performed for a selected case as well (\cite{Schaab98c}). As far as we know, this is the first fully two-dimensional simulation of neutron star cooling which takes internal heating into account. The heating processes emerge as a response to the loss of rotational energy of the star and can be divided roughly into two groups. Firstly, processes caused by the readjustment of the equilibrium structure of the compact object such as a sequence of crust/core-quakes (\cite{Cheng92a} and references therein); nuclear reactions in cold nuclear matter with non-equilibrium composition (picno-nuclear reactions in the crust and weak processes in the core, see, e.g., \cite{Iida96a} and \cite{Reisenegger92a}). Secondly, processes related to superfluidity such as dissipative motion of the neutron vortex lattice in the superfluid crusts (\cite{Shibazaki89}; \cite{Link91a}; \cite{Bildsten89a}; \cite{Jones90a}) and the dissipative motion of proton vortex lattice in the superfluid core and the vorticity decay at the phase separation boundaries, e.g. at the crust-core interface, (\cite{Sedrakian93}; \cite{Sedrakian98a}). The rate of dissipation associated with these processes is, in most cases, intimately connected with the spin evolution of the star, in particular the evolution of its magnetic moment. The time dependence of the magnetic moment and the moment of inertia will affect the heating rate and thus the star's cooling behaviour along with the relation between the luminosity and the age of the star; these effects are ignored in this study in order to constrain the actual number of the evolutionary scenarios; we shall assume therefore a constant magnetic moment and moment of inertia. The paper is organised as follows. The equations of thermal and rotational evolution of a neutron star are given in Sect. \ref{sec:equations}. Various heating processes associated with the dissipative motion of vortices and the changes of the equilibrium structure are reviewed in Sects. \ref{sec:vortex} and \ref{sec:nonequilibrium}. Additional physical input is discussed in Sect. \ref{sec:inp}. The observed data are summarised in Sect. \ref{sec:observations}. Section \ref{sec:res} presents the results of our cooling simulations which are discussed in Sect. \ref{sec:disc}. \section{Thermal and rotational evolutions} \label{sec:equations} \subsection{Equations of thermal evolution} \label{sec:eq.thermal} The thermal evolution is governed by the coupled system of equations for energy balance (\cite{Thorne77}), \begin{align} \frac{\partial(L{\rm e}^{2\phi})}{\partial r} &= 4\pi r^2 {\rm e}^\Lambda\left(-\epsilon_\nu{\rm e}^{2\phi} +h{\rm e}^{2\phi} -c_\mathrm{v}\frac{\partial(T{\rm e}^\phi)}{\partial t} \right)\, , \label{eq:ebal} \\ \intertext{and thermal energy transport,} \frac{\partial(T{\rm e}^{\phi})}{\partial r} &= - \frac{(L{\rm e}^{2\phi}){\rm e}^{\Lambda-\phi}}{4\pi r^2\kappa} \, . \label{eq:etran} \end{align} This system requires as a microphysical input the neutrino emissivity per unit volume, $\epsilon_\nu(\rho,T)$, the heat rate per unit volume, $h(\rho,T,\Omega,\dot\Omega)$, the heat capacity per unit volume, $c_\mathrm{v}(\rho,T)$, and the thermal conductivity, $\kappa(\rho,T)$, where $\rho$ is the local density. The $g_{tt}$ and $g_{rr}$ components of the Schwarzschild metric tensor $\bf g$ are denoted by $-{\rm e}^{2\phi}$ and ${\rm e}^{2\Lambda}$, respectively, $L$ is the luminosity, and the other variables have their usual meaning. The boundary conditions for (\ref{eq:ebal}) and (\ref{eq:etran}) read \begin{align} L(r=0) &= 0 \, , \label{eq:bdl} \\ T(r=r_\mathrm{m}) &= T_\mathrm{m}(r_\mathrm{m},L_\mathrm{m},M_\mathrm{m}) \, , \label{eq:bdt} \end{align} where $T_\mathrm{m}(r_\mathrm{m},L_\mathrm{m},M_\mathrm{m})$ is fixed by the properties of the photosphere at the density $\rho=\rho_\mathrm{m}=10^{10}\mathrm{\,g\,cm}^{-3}$ (\cite{Gudmundson83}). Since the star's thermal evolution for times greater than, say, a few months does not depend on the detailed initial temperature profile in the interior of the star, one can choose, without loss of generality, an initialy isothermal temperature distribution of $T(r)\equiv 10^{11}$~K. The set of equations for thermal evolution (\ref{eq:ebal}) and (\ref{eq:etran}) were solved numerically by means of a Newton-Raphson type algorithm. The microphysical input quantities needed for the solution of these equations are given in Table \ref{tab:ingredients} and will be discussed in detail in Sect. \ref{sec:inp}. If one accounts for azimuthal asymmetry of the input quantities, e.g. the heat rate $h$, one has to solve two-dimensional thermal evolution equations, which have already been presented in Schaab \& Weigel (1998) \nocite*{Schaab98c} along with the numerical method for solving them. \subsection{The normal component} \label{sec:spin.normal} We shall start the discussion of the rotational evolution by a brief review of the spin dynamics of the normal (non-superfluid) component of the star. The combined effect of the braking of the star's rotation via the emission of electromagnetic radiation, electron-positron wind, and gravitational waves on the spin evolution is commonly cast in a generic power-law form \begin{equation} \label{eq:rot.evol} \dot\Omega_{\rm c}(t) = -K(t)\Omega_{\rm c}^n(t), \end{equation} where $\Omega_{\rm c}$ is the spin frequency of the normal component. The functional form of $K(t)$ and $n(t)$ depends on the dominant process assumed. If one assumes that the energy losses are due to the magnetic dipole radiation (\cite{Goldreich71a}), then $n = 3$ and $K$ is a function of the surface magnetic field strength $B$, the angle $\alpha$ between spin and magnetic axis, the moment of inertia of the star $I$, and a correction factor $\gamma\equiv 1+{\rm d}\ln I/ {\rm d}\ln\Omega_{\rm c}$: \begin{equation} K \propto \frac{B^2 \sin^2\alpha}{\gamma I}~. \end{equation} The braking index is known for the four youngest pulsars for which the measurements of the second time derivative of the period, $P$, are available (\cite{Lyne93a,Kaspi94a,Boyd95a}). All four braking indices given by $n= P\ddot P /\dot P^2$ fall into the range between 1.2 and 2.8. The deviations from the value for the magnetic dipole radiation has been attributed to the increase of the surface magnetic field strength (\cite{Blandford83a,Muslimov96a}), an increase of the angle between the spin and magnetic moment axis (\cite{Beskin84a,Link97a}), or presence of the weakly coupled superfluid in the core of the star (\cite{Sedrakian98b}). We shall assume, in a first approach to the problem, $K$ to be constant and $n = 3$. With these assumptions, integration of Eq. \eqref{eq:rot.evol} yields \begin{equation} \label{eq:rot.omega} \Omega_{\rm c}(t) = \left( (n-1)Kt+\Omega_{\rm i}^{-(n-1)}\right) ^{-1/(n-1)}, \end{equation} where $\Omega_{\rm i}$ is the initial angular velocity at $t=0$. If the initial angular velocity is large compared to the present angular velocity, Eq. \eqref{eq:rot.omega} simplifies to \begin{equation} \Omega_{\rm c}(t) = \left( (n-1)Kt \right)^{-1/(n-1)}, \end{equation} and the age of the pulsar can be expressed as \begin{equation} t = \frac{1}{n-1}\frac{\Omega_{\rm c}(t)}{|\dot\Omega_{\rm c}(t)|}. \end{equation} Apart from affecting the cooling history, the time variation of $K$, as well as a deviation of the braking index from its canonical value, $n=3$, may lead to over- or underestimating the true age by a factor as large as 3. A comparison of the theoretical cooling tracks with observed data is therefore affected by this uncertainty (see Sect. \ref{sec:observations}). \subsection{The superfluid component} \label{sec:spin.sf} The angular velocity of the superfluid phases adjusts to the changes in the rotation rate of the normal component via expansion or contraction of the neutron vortex lattice, which carries the angular momentum of the superfluid. The mean macroscopic density of neutron vortex lines $n_{\rm l}$ in cylindrical coordinates $(r_{\rm p},z,\phi)$ is given by the Feynman-Onsager quantisation condition \begin{equation} \label{eq:vortex.dens} \kappa n_{\rm l}(r_{\rm p},t) = 2\Omega_{\rm s}(r_{\rm p},t)+r_{\rm p} \frac{\partial\Omega_{\rm s}(r_{\rm p},t)}{\partial r_{\rm p}}. \end{equation} $\Omega_{\rm s}(r_{\rm p})$ denotes the angular velocity of the superfluid component and $\kappa = h/2m_{\rm N}$ is the quantum of circulation carried by each vortex, where $h$ denotes the Planck constant and $m_{\rm N}$ is the bare neutron mass. The second term accounts for the gradients in the distribution of vortex lines. Because of the lack of sources and sinks for vortex lines in the bulk of the superfluid, the temporal evolution of the vortex number density obeys the number conservation law: \begin{equation}\label{eq:conserv} \frac{\partial n_{l}}{\partial t} +{\vec\nabla}\cdot(n_{ l}{\vec v_{ l}}) = 0, \end{equation} where ${\vec v_{l}}$ is the velocity of vortex lines. Combination of Eqs. \eqref{eq:vortex.dens} and \eqref{eq:conserv} leads to \begin{equation} \label{eq:creeprate} \frac{\partial\Omega_{\rm s}}{\partial t} = -\frac{1}{r_{\rm p}}\kappa n_{\rm l}v^{\rm r}_{\rm l} = -\left( \frac{2}{r_{\rm p}}\Omega_{\rm s} +\frac{\partial\Omega_{\rm s}}{\partial r_{\rm p}} \right) v^{\rm r}_{\rm l}. \end{equation} A decrease of the spin of the superfluid component ($\dot\Omega_{\rm s}<0$) causes an outward radial velocity $v^{\rm r}_{\rm l}>0$. In the following we neglect the second term in parentheses and assume $\Omega_{\rm s}(r_{\rm p})$ to be constant throughout the superfluid phase, since the vorticity gradients depend on the density profile of the mutual friction coefficients, which are poorly known. Furthermore, over the evolutionary timescale one can assume that $\dot \Omega_c = \dot \Omega_{\rm s}$, i.e. the deceleration rates of the superfluid and the normal components are the same. Then the velocity of the radial motion of the vortices (and therefore the rate of dissipation) are related to the spin-down characteristics of the normal component via Eq. \eqref{eq:creeprate}. In this manner the dynamics of superfluid phases is eliminated from the system of coupled equations of secular and thermal evolution, and the heating rates are given directly in terms the spin characteristics of the normal component. \section{Dissipative motion of vortex lattices} \label{sec:vortex} We next turn to the description of the dissipation processes included in the thermal evolution simulations. A number of dynamical coupling regimes has been suggested in the literature for the coupling of the superfluid component to the crusts. We shall discuss several variants of the two main alternatives - the pinned and the corotation regimes. \subsection{Vortex pinning at nuclei} \label{sec:pinning} The vortex creep theories of the dynamics of neutron vortices in the inner crust assume these vortices to be pinned to the nuclear lattice (\cite{Anderson75a,Alpar77a,Alpar84,Epstein88a,Link91a,Pizzochero97a}). The pinning force provides a large resistive force to the vortex motion which is necessary to spin down the superfluid component. In this regime, the vortices move due to the process of vortex creep, i.e. by thermal activation from one pinning configuration to another. The vortex creep rate has the general form (\cite{Link93}) \begin{equation} \label{eq:micro.creeprate} v^{\rm r}_{\rm l} = v_0 {\rm e}^{-A/T_{\rm eff}}, \end{equation} where $v_0$ is a microscopic velocity, $A$ denotes the activation energy for a segment of a vortex line to overcome its pinning barrier, and $T_{\rm eff}$ is the effective temperature which includes quantum tunnelling effects and is a function of the local temperature of the thermal bath. These quantities depend on microscopic parameters (e.g. the pinning energy), which themselves depend on the density, and on the velocity difference between the superfluid and the normal component $\delta v^\phi$. Since the creep rate $v^{\rm r}_{\rm l}$ is given by Eq. \eqref{eq:creeprate}, Eq. \eqref{eq:micro.creeprate} has to be inverted to obtain the velocity difference $\delta v^\phi$. This is done by an iterative procedure similar to the one described by Van Riper et al. (1995) \nocite{VanRiper95a}. The rotational energy of the superfluid is converted to thermal energy by the radial motion of the vortices. The energy dissipated by a single vortex per unit time is $\dot E_{\rm diss}=f_{\rm M}^{\rm r}v_{\rm l}^{\rm r}$, where $f_{\rm M}^{\rm r}=\rho_{\rm s}\kappa\delta v^\phi$ is the radial component of the Magnus force. The dissipated energy $h$ per unit time and volume is then given by: \begin{equation} \label{eq:diss.pin} h = \dot E_{\rm diss}n_{\rm l} = r_{\rm p}\rho_{\rm s}\delta v^\phi |\dot\Omega_{\rm s}|, \end{equation} where Eq. \eqref{eq:creeprate} has been used. The integration over the volume of the superfluid component gives: \begin{equation} \label{eq:pinn.heatrate} \int h{\rm d} V = |\dot\Omega_{\rm s}| \int \rho_{\rm s}r_{\rm p}\delta v^\phi {\rm d} V. \end{equation} This result guarantees conservation of both energy and angular momentum. For the implementation of $h$ in our one-dimensional cooling code, $h$ has to be averaged over spherical shells: \begin{equation} \bar h =\frac{\int h \sin\theta {\rm d}\theta{\rm d}\phi}{4\pi} = \frac{1}{4}\pi r\rho_{\rm s}\delta v^\phi |\dot\Omega_{\rm s}|, \end{equation} where we assumed that $\delta v^\phi$ is approximately constant on spherical shells. This approximation seems to be reliable because $\delta v^\phi$ depends only logarithmically on $r_{\rm p}$ (\cite{VanRiper95a}). We shall consider two models for pinning: the first model (EB) uses the parameters from Epstein \& Baym (1988) \nocite{Epstein88a} and Link \& Epstein (1991) \nocite{Link91a}, where the pinning energy is derived from the Ginzburg-Landau theory. In the second model (PVB) the parameters are derived by Pizzochero et al. (1997) \nocite{Pizzochero97a} from the Bardeen-Cooper-Schrieffer theory of superconductivity. Below the density $1.5\times10^{13}\mathrm{\,g\,cm}^{-3}$ the parameters for the both models do not deviate considerably. However, above densities $\gtrsim 1.5\times10^{13}\mathrm{\,g\,cm}^{-3}$ the large deviations in the pinning energy (the typical pinning energy is $\sim 10{\rm\,MeV}$ in the EB model and $\sim 1{\rm\,MeV}$ in the PVB model) leads to rather different evolutionary scenarios and, therefore, we shall treat both models seperately. \subsection{Corotating regime in the crust} \label{sec:dragcrust} The dynamics of the neutron vortices in the corotating regime is governed by the force balance equation \begin{equation} {\vec f}_{\rm M}+{\vec f}_{\rm d} = 0, \end{equation} where the Magnus and the drag (resistive) forces are: \begin{align} {\vec f}_{\rm M} &= -\rho_s {\vec\kappa}\times ({\vec v}_{\rm s}-{\vec v}_{\rm l}), \\ {\vec f}_{\rm d} &= -\eta ({\vec v}_{\rm l}-{\vec v}_{\rm c}), \end{align} with the drag coefficient $\eta$. A decomposition of the force balance equation into radial and azimuthal components gives \begin{equation} {\vec v}_{\rm l} = (v_{\rm s}-v_{\rm c})\sin\theta_{\rm d} \cos\theta_{\rm d}{\vec e}_r +(v_{\rm s}+v_{\rm c}\tan^2\theta_{\rm d})\cos^2\theta_{\rm d} {\vec e}_\phi, \end{equation} where the dissipation angle $\theta_{\rm d}$, following Bildsten \& Epstein (1989)\nocite{Bildsten89a}, is defined as \begin{equation} \tan\theta_{\rm d} = \frac{\eta}{\rho_{\rm s}\kappa}. \end{equation} The rate of the dissipation per unit volume is then given by \begin{equation} h = -{\vec f}_{\rm d}\cdot({\vec v}_{\rm l}-{\vec v}_{\rm c}) n_{\rm free} = \rho_{\rm s}r_{\rm p}^2|\dot\Omega_{\rm s}|\omega, \end{equation} with the abbreviation \begin{align} \omega &= \frac{v_{\rm s}-v_{\rm c}}{r} = \frac{v^{\rm r}_{\rm l}} {r_{\rm p}\sin\theta_{\rm d}\cos\theta_{\rm d}}, \\ &= -\frac{\dot\Omega_{\rm s}}{2\Omega_{\rm s}}\frac{1}{\Phi_{\rm free}} \frac{1}{\sin\theta_{\rm d}\cos\theta_{\rm d}}. \end{align} Here, $\Phi_{\rm free}$ denotes the fraction of the free (corotating) vortices. This quantity is largely unknown, but for stationary situations, one may assume that definite crustal regions support either corotation or creep regimes and the mixing is small. Hence we shall adopt $\Phi_{\rm free}:=n_{\rm free}/n_{\rm l}=1$ for the corotating regime. The dominant processes in this regime result from the coupling of the translational motion of the vortices to the electron-phonon system (\cite{Jones90a}). The leading order coupling implies a resistive force constant \begin{equation}\label{ETA0} \eta_0 \simeq \frac{3}{32\pi^{1/2}}\frac{a \,E_p^2}{M\xi_n^3c_s^3}, \end{equation} where $a$ is the lattice constant, $E_p$ the pinning energy, $\xi_n$ the neutron coherence length, $M$ the (effective) mass of the nuclei, and $c_s$ the phonon velocity (given by the unperturbed dispersion relation). Using eq. (\ref{ETA0}) we find $\eta_0 \sim 10^5$ g cm$^{-1}$ s$^{-1}$ at the density $\rho_s = 10^{13}$ g cm$^{-3}$ and, therefore, $\tan\theta_d \sim 5 \times 10^{-6}$. The processes corresponding to higher orders in the vortex displacement are generally velocity dependent. The next-to-leading order term, involving one-phonon processes, is estimated as (\cite{Epstein92a,Jones92a}) \begin{equation} \eta_1 \simeq \frac{E_p^2} {2\rho_s\omega a\xi_n^2\sqrt{2\pi\xi_n c_K v_L^3}}, \end{equation} where $c_K$ is the kelvon velocity due to excitation of the oscillatory degrees of freedom of the vortex lines moving with velocity $v_L$. This process is sensitive to the actual excitation spectrum of kelvons. The possible localisation of kelvons in a random nuclear potential (\cite{Jones92a}) leads to a threshold in the excitation spectrum beyond which a continuum of states becomes available for excitations. In addition one requires knowledge of the scattering amplitudes beyond the Born approximation. For typical parameters one finds $\eta_1 \sim 10^{10}$ g cm$^{-1}$ s$^{-1}$ at a density of $\rho_s\sim 10^{13}$ g cm$^{-3}$, and, therefore, $\tan\theta_d \sim 1$. However, in the stationary situations, the vortex velocities are rather small, and we shall assume that the leading term $\eta_0$ is the dominant one in the corotating regime. \subsection{Electron--vortex scattering in superfluid core} \label{sec:electronscatt} An estimate of the heating processes in the superfluid core requires knowledge of vortex lattice structure that nucleates after the superfluid phase transition. We shall pursue the point of view that the core's magnetic field nucleates in a proton vortex lattice. The frictional forces in a neutron star's core are dominated by the scattering of the quasiparticle excitations of the electron normal Fermi-liquid by the proton vortex lattice. The reason is that the number of proton vortices per neutron vortex, given by $n\simeq (\sqrt{3}/2) (B/\Phi_0) d_n^2$, is of the order of $n\ge 10^{12}$ for $B\sim 10^{12}$ G and a neutron vortex lattice constant of $d_n \sim 10^{-3}$ cm. The dominant dissipative scattering results from the scattering of electron excitations off the quasiparticle excitations confined in the core of proton vortex lines. The scattering rate is a sensitive function of the form-factor of the vortex line core and the actual distribution of vortex lines (e.g. formation of vortex clusters, \cite{Sedrakian98a}). An order-of-magnitude estimate of the dissipation rate is given by (\cite{Sedrakian93}) \begin{equation} h = \frac{45}{8\pi} f T \left(\frac{\dot\Omega}{\Omega}\right)^2R^2\sin^2\theta \cos^2\theta, \end{equation} where $T$ is the temperature, $R$ the radius of the superfluid core and $f$ is defined as\footnote{Note that the multiplication sign in the second line of Eq. (11) in Sedrakian \& Sedrakian (1993)\nocite{Sedrakian93} should be replaced by a division sign. This misprint is corrected here.}. \begin{multline} f = 41\pi^2\alpha^2\frac{m_{\rm p}^2k_{\rm B}}{\hbar^3c} \frac{k_{\rm F}}{k_{\rm FT}}\frac{\Delta_{\rm p}E_{\rm p}}{\hbar} |k|\left(\frac{\xi_{\rm p}}{\lambda_{\rm p}}\right)^{2+2/3|k|} \\ \times\left(\frac{2m_{\rm p}c^2}{E_{\rm p}}\right)^{1/2} \exp\left(-\frac{0.78\Delta_{\rm p}^2}{E_{\rm p}k_{\rm B}T}\right), \end{multline} where $k_{\rm F} = (3\pi^2n_{\rm p})^{1/3}$ denotes the proton Fermi momentum, $k_{\rm FT} = \left(4k_{\rm F}m_{\rm p}e^2/\pi\hbar^2\right)^{1/2}$ the Thomas-Fermi screening length, $E_{\rm p} = \hbar^2k_{\rm F}^2/ {2m_{\rm p}^\ast}$ the proton quasiparticle energy with the proton effective mass $m_{\rm p}^\ast$, $ k \equiv (m_{\rm p}^\ast-m_{\rm p})/{m_{\rm p}}$ the entrainment coefficient, and \begin{align} \xi_{\rm p} &= \frac{\hbar^2k_{\rm F}}{\pi m_{\rm p}\Delta_{\rm p}}, \\ \lambda_{\rm p} &= (1+|k|)^{1/2} \left(\frac{m_{\rm p}c^2}{4\pi n_{\rm p}e^2}\right)^{1/2} \end{align} the proton coherence length and the magnetic field penetration length, respectively. It should be noted that the dynamical coupling of the proton vortex clusters to the electron liquid is controlled by the electron flux-scattering processes, whose rate is larger than that for the electron scattering off the vortex core quasiparticles. While in the latter case the scattering is inelastic (i.e. energy is transfered to the proton quasiparticle localised in the vortex core), the former process of electron-flux scattering is predominantly elastic i.e. involves only a change of direction of the electron momentum (but no changes in the energy). In the classical (or quasi-classical) limit this process corresponds to the bending of the electron trajectory in the magnetic field of a vortex under the action of the Lorentz force and is manifestly non-dissipative. In the quantum limit the processes of the soft-photon emission can not be dismissed in general; however the corresponding rates should be substantially less than the elastric scattering rates. \subsection{Decay of vortices} \label{sec:vortexdecay} The proton vortices, which are dragged along by the neutron vortex motion in the radial outward direction to the crust--core boundary, would decay as they merge in the crust, thus releasing their self-energy. The relevant crust-core boundary for this process should be identified with the phase transition point between the phase with protons bound in heavy nuclei and the phase in which they are in continuum states. Since the crustal matter can not support a proton supercurrent, the circulation currents of proton vortices would decay on the ohmic dissipation timescale, which is much smaller than a star's spin-down time scale. The resulting heating rate at the crust-core boundary is given by \begin{equation} \int h{\rm d} V = g \frac{|\dot\Omega|}{\Omega}R^3, \end{equation} with \begin{multline} g = \frac{1}{3} \left(1-\left(2\ln\frac{\lambda_{\rm p}}{\xi_{\rm p}}\right)^{- 1}\right)|k| \left(\frac{\xi_{\rm p}}{\lambda_{\rm p}}\right)^{2/3|k|} \\ \times\left(\frac{\Phi_0}{4\pi\lambda_{\rm p}}\right)^2 \ln\frac{\lambda_{\rm p}}{\xi_{\rm p}} \, . \end{multline} Here, $\Phi_0 = {hc}/{2e}$ is the flux quantum. The quantity $g$ is very sensitive to variations of the microphysical parameters such as the proton density, the effective mass, and the gap energy at the location of the phase transition point. We use a fixed upper limit for this process corresponding to the value of $g= 10^{28}$. \section{Readjustment to equilibrium structure} \label{sec:nonequilibrium} As a neutron star spins down its global structure readjusts itself toward a more spherically symmetric configuration to minimise the sum of gravitational and rotational energy. Two models of internal heating discussed here are based on the structural readjustment of the oblateness of the star, defined as \begin{equation} e_{\rm eq} := \frac{I(\Omega)}{I(0)}-1, \end{equation} and of the central baryon density $n_{\rm c}$. $I(\Omega)$ denotes the star's moment of inertia. The equilibrium values of both quantities are determined as functions of the angular velocity by computing a sequence of rotating neutron star models with constant baryon number (\cite{Schaab97a}). These calculations are done by solving the general relativistic stellar structure equations in two dimensions, in a similar manner as in Komatsu et al. (1989)\nocite{Komatsu89a} and Bonazzola et al. (1993)\nocite{Bonazzola93a}. As an example the oblateness $e$ and the fraction $1-n_{\rm c}(\Omega)/n_{\rm c}(0)$ with the central density $n_{\rm c}(\Omega)$ are plotted for a neutron star model with constant baryon mass $M_{\rm B}=1.58M_\odot$ in Fig. \ref{fig:rot}. The underlying equation of state is the non-relativistic UV$_{14}$+UVII -model (see Sect. \ref{sec:inp.eos}). \begin{figure} \resizebox{\hsize}{!}{\rotatebox{-90}{\includegraphics{h1252.f1}}} \caption{Oblateness $e$ and the variation of the central density $1-n_{\rm c}(\Omega)/n_{\rm c}(0)$ of rotating neutron stars versus angular velocity. The parameters are: $M_{\rm B}=1.58M_{\odot}$ and UV$_{14}$+UVII -EOS.} \label{fig:rot} \end{figure} \subsection{Crust cracking} \label{sec:noneq.cracking} If both the interior and the crust are liquid, the oblateness will continuously adjusts itself to the equilibrium value. This is the case for the temperatures $T\gtrsim 10^8$~K (\cite{Cheng92a}). However, if the crust solidifies, the oblateness continuously depart from its equilibrium value causing an increasing mean stress on the crust according to (\cite{Baym71c}) \begin{equation} \sigma = \tilde{\mu}(e-e_{\rm eq}), \end{equation} where $\tilde{\mu}$ is the mean shear modulus of the crust. The quantity $e_{\rm eq}$ refers to the equilibrium value of the oblateness. If the mean stress $\sigma$ reaches some critical value \begin{equation} \label{eq:maxshearangle} \sigma_{\rm c} = \tilde{\mu}\theta_{\rm c}, \end{equation} the crust breaks and the deviation of $e$ from its equilibrium value is reduced. For an ideal Coulomb lattice, the value of the critical shear angle $\theta_{\rm c}$ is of the order of magnitude $10^{-1}-10^{-2}$ (\cite{Smolukowski70a}). The fraction of impurities or defects in the neutron star crusts could be considerable due to, e.g., the fast cooling below the melting temperature of the lattice. Smolukowski (1970)\nocite{Smolukowski70a} argues that the value of $\theta_{\rm c}$ is then reduced to roughly $10^{-4}-10^{-5}$. During the breaking of the crust the strain energy \begin{equation} \Delta E_{\rm strain} = 2B(e-e_{\rm eq})\Delta(e-e_{\rm eq}) \end{equation} is released (\cite{Cheng92a}), where $\Delta(e-e_{\rm eq})\ll(e-e_{\rm eq})$ is the change in $(e-e_{\rm eq})$, and the coefficient $B$ is given in terms of the mean shear modulus by $B=\frac{1}{2}V_{\rm c}\tilde{\mu}$ with $V_{\rm c}$ being the volume of the crust. We follow Cheng et al. (1992)\nocite{Cheng92a} in assuming that $(e-e_{\rm eq})$ is approximately equal to its maximum value $\theta_{\rm c}$ and that the time $\Delta t$ between two successive quakes is small compared to the time scale of thermal evolution. Hence\footnote{Our formulas differ from those of Cheng et al. (1992)\nocite{Cheng92a} by a factor $\theta_{\rm c}/e$, which is $\ll 1$ for slowly rotating stars. We therefore obtain no effect on the cooling of slowly rotating stars, in contrast to Cheng et al. (1992)\nocite{Cheng92a} where rather large effects are claimed.} \begin{equation} \label{eq:strainenergy} \frac{{\rm d} E_{\rm strain}}{{\rm d} t} \approx \frac{\Delta E_{\rm strain}}{\Delta t} = 2B\theta_{\rm c}\frac{\Delta(e-e_{\rm eq})}{\Delta t} \approx -2B\theta_{\rm c}\frac{{\rm d} e_{\rm eq}}{{\rm d}\Omega_{\rm c}} \dot\Omega_{\rm c}. \end{equation} The crust strength $B$ is given by the integral (\cite{Baym71c}) \begin{equation} B = \int_{V_{\rm c}} b(r){\rm d}^3r \end{equation} with \begin{equation} b(r) = \frac{1}{25} C_{44}(r)\left(96-166\left(\frac{r}{R}\right)^2 +\frac{777}{10}\left(\frac{r}{R}\right)^4\right) \end{equation} and the shear modulus of a bcc-lattice (\cite{Mott36a}) \begin{equation} C_{44}(r) = 0.3711 Z^2e^2n_{\rm N} \left(\frac{2}{n_{\rm N}}\right)^{-1/3}. \end{equation} The total strain energy release \eqref{eq:strainenergy} can thus be expressed in terms of a local energy loss rate: \begin{equation} h = 2b\theta_{\rm c}\frac{{\rm d} e_{\rm eq}}{{\rm d}\Omega_{\rm c}} |\dot\Omega_{\rm c}|. \end{equation} \subsection{Chemical heating} \label{sec:chem} The central density of a rotating neutron star with fixed baryon number depends on the angular velocity (see Fig. \ref{fig:rot}). As the star spins down, the centrifugal force decreases and correspondingly the central density increases (by up to 20~\%). This is accompanied by a shift of the chemical equilibrium. The matter would maintain chemical equilibrium if the relaxation timescales for the weak processes are small compared to the timescale of rotational evolution. However, these timescales are found to be comparable and the actual composition departs from chemical equilibrium, which modifies the reaction rates (\cite{Haensel92a}) and usually leads to a net conversion of chemical energy into thermal energy (\cite{Reisenegger92a,Iida96a}). For simplicity we restrict ourselves to a system containing only neutrons, protons, electrons, and muons in $\beta$-equilibrium (as is described by the non-relativistic EOS UV$_{14}$+UVII). The equation of state of cold, charge neutral matter, which is not necessarily in $\beta$-equilibrium (we assume however equal chemical potentials for electrons and muons: $\mu_{\rm e}=\mu_\mu$.), therefore depends only on two parameters. These are chosen to be the baryon density $n$ and the relative proton density $x=n_{\rm p}/n$. These two parameters are constant on closed, sphere-like surfaces $\mathcal{S}(N)$ of constant energy density $\rho$ and pressure $P$. $N$ refers to the number of baryons enclosed by $\mathcal{S}$. We approximate the time derivative of the baryon density on a specific surface $\mathcal{S}(N)$ by the expression (\cite{Reisenegger92a}) \begin{equation} \label{eq:dens.var} \frac{{\rm d} n}{{\rm d} t} \simeq \left. \frac{n}{n_{\rm c}} \right\vert_{\Omega=0} \frac{{\rm d} n_{\rm c}}{{\rm d}\Omega} \frac{{\rm d}\Omega}{{\rm d} t}, \end{equation} where $n_{\rm c}$ is the central baryon density, and ${\rm d} n_{\rm c} / {\rm d}\Omega$ is obtained from calculations of rotating neutron star sequences (see Fig. \ref{fig:rot}). The equilibrium value $x_{\rm eq}$ of the proton fraction is determined by the following condition on the energy per baryon $E$: \begin{align} \delta\mu := \frac{\partial E}{\partial x} &= \frac{\partial E}{\partial n_{\rm p}}\frac{\partial n_{\rm p}}{\partial x} +\frac{\partial E}{\partial n_{\rm n}}\frac{\partial n_{\rm n}}{\partial x} +\frac{\partial E}{\partial n_{\rm e}}\frac{\partial n_{\rm e}}{\partial x} +\frac{\partial E}{\partial n_\mu}\frac{\partial n_\mu}{\partial x} \\ &= \mu_{\rm p}-\mu_{\rm n}+\mu_{\rm e} = 0, \end{align} where we used the charge neutrality of the system, $\partial n_{\rm p}/ \partial x = \partial n_{\rm e}/ \partial x + \partial n_\mu/ \partial x$, the condition $\mu_{\rm e}=\mu_\mu$, and the equation $\partial n_{\rm p}/ \partial x=-\partial n_{\rm n}/ \partial x=n$. The time derivative of $n$ is linked to the time derivative of $x^{\rm eq}$ by \begin{equation} \frac{{\rm d} x^{\rm eq}}{{\rm d} t} \simeq \frac{{\rm d} x^{\rm eq}}{{\rm d} n} \left.\frac{n}{n_{\rm c}}\right|_{\Omega=0} \frac{{\rm d} n}{{\rm d}\Omega} \frac{{\rm d}\Omega}{{\rm d} t}, \end{equation} where Eq. \eqref{eq:dens.var} was used. A deviation from chemical equilibrium corresponds to a non-vanishing value of $\delta\mu$. The non-vanishing gradient of the total chemical potential including both internal and external contributions causes a diffusion of particles. If the timescale for chemical relaxation is larger than the timescale for this diffusion, the chemical relaxation takes place in a larger regoin of the neutron star, and $\delta\mu$ has to be averaged over such regions. It seems likely that this kind of relaxation occurs in the core of neutron stars (\cite{Reisenegger96a}). The time derivative of \begin{equation} \delta\mu = \frac{\partial E}{\partial x} = \left. \frac{\partial^2E}{\partial x^2}\right\vert_{x=x_{\rm eq}} (x-x_{\rm eq}) \end{equation} yields the evolution equation for $\delta\mu$: \begin{equation} \frac{\partial\delta\mu}{\partial t} = \left. \frac{\partial^2E}{\partial x^2}\right\vert_{x=x_{\rm eq}} \left( \frac{\Gamma}{n}-\frac{{\rm d} x^{\rm eq}}{{\rm d} n} \left.\frac{n}{n_{\rm c}}\right|_{\Omega=0} \frac{{\rm d} n}{{\rm d}\Omega} \frac{{\rm d}\Omega}{{\rm d} t}\right), \end{equation} where $\Gamma(\delta\mu,n)=n\partial x/ \partial t$ is the difference between the rates per unit volume of electron capture and beta decays. This difference gives a positive value for $\delta\mu>0$, so that chemical equilibrium tends to be restored. The second derivative of the energy per baryon is equal to \begin{equation} \left.\frac{\partial^2E}{\partial x^2}\right\vert_{x=x_{\rm eq}} = 8S(n)+\frac{1}{3}\hbar c(3\pi n)^{1/3}x^{-2/3}, \end{equation} where $S(n)$ denotes the symmetry energy. The average energy released by these reactions into kinetic and thus thermal energy is equal to the difference of the Fermi energies of protons and electrons on the one side and neutrons on the other side, i.e.: \begin{equation} h = \Gamma\delta\mu\, . \end{equation} At the same time, thermal energy is lost by emission of thermal neutrinos, as it is the case for matter in chemical equilibrium. The emission rates are however modified by the deviation from chemical equilibrium (see \cite{Reisenegger92a} for the values of $\Gamma$ and $\epsilon_\nu$). Like the neutrino emissivity $\epsilon_\nu$, the reaction rate $\Gamma$ is suppressed by superfluidity unless the deviation of chemical equilibrium $\delta\mu$ exceeds the sum of the gap energies of the participating baryons (\cite{Reisenegger96a}). Until now we have considered only the restoration of chemical equilibrium in the core. The situation in the crust is similar, though more complex. Iida \& Sato (1997)\nocite{Iida96a} studied the Lagrangian changes in pressure associated with elements of matter due to the spin-down in the framework of the Hartle-approximation (\cite{Hartle67a}). By considering the nuclear processes induced above $\rho_{\rm l}=3\times 10^{13}\mathrm{\,g\,cm}^{-3}$ by compression or decompression of matter, they approximate the chemical energy per unit volume and time converted to thermal energy to \begin{equation} h \approx nqN\sin\theta|1-1.46\cos^2\theta|\frac{2\Omega|\dot\Omega|}{(6283{\rm\, s}^{-1})^2}\, , \end{equation} where $nq\approx 4{\rm\, eV\, fm}^{-3}$ is the energy per unit volume released by one non-equilibrium process, $N\approx 15(\rho-\rho_{\rm l})/(\rho_{\rm tr}-\rho_{\rm l})$ is the number of processes in a unit cell, and $\rho_{\rm tr}=1.7\times 10^{14}\mathrm{\,g\,cm}^{-3}$ is the transition density between core and inner crust. By averaging over spherical shells we obtain \begin{equation} h \approx 3\times 10^{21} \frac{\rho-\rho_{\rm l}}{\rho_{\rm tr}-\rho_{\rm l}} \frac{\Omega|\dot\Omega|}{1{\rm\, s}^{-3}} {\rm\, erg\, cm}^{-3}{\rm\,s}^{-1} \, . \end{equation} \section{Additional ingredients and comparison of the efficiency of the heating mechanisms} \label{sec:inp} Apart from the rates of the various heating processes discussed in the previous sections, some further ingredients are needed to solve the general relativistic equations of stellar structure and evolution (see Sect. \ref{sec:eq.thermal}). These additional ingredients are summarised in Table \ref{tab:ingredients}. Here we shall only discuss the equation of state of the dense interior of a neutron star, the neutrino emissivities, the superfluidity gaps and the photosphere, since these ingredients are the most important ones. Further, we will compare the efficiencies of the various heating mechanism. \begin{table*} \centering \caption[]{Input quantities used for the cooling simulations \label{tab:ingredients}} \begin{tabular}{ll} \hline\noalign{\smallskip} Parameter & References \\ \noalign{\smallskip}\hline\noalign{\smallskip} Equations of state: \\ \quad crust & \cite{Haensel94a}, \cite{Negele73} \\ \quad core (alternatives) & \\ \qquad UV$_{14}$+UVII & \cite{Wiringa88} \\ \qquad RHF8 & \cite{Huber97a} \\ \hline Superfluidity & see Table \ref{tab:sf} \\ \hline Heat capacity & \cite{Shapiro83}, \cite{VanRiper91} \\ \hline Thermal conductivities: \\ \quad crust & \cite{Itoh84a}, \cite{Itoh83a}, \cite{Mitake84} \\ \quad core & \cite{Gnedin95a} \\ \hline Neutrino emissivities: \\ \quad pair-, photon-, plasma-processes & \cite{Itoh89} \\ \quad bremsstrahlung in the crust & \cite{Yakovlev96a}, \cite{Haensel96a} \\ \quad bremsstrahlung in the core & \cite{Friman79}, \cite{Kaminker97a} \\ \quad modified Urca & \cite{Friman79}, \\ & \cite{Yakovlev95a} \\ \quad direct nucleon Urca & \cite{Lattimer91} \\ \quad direct hyperon Urca & \cite{Prakash92} \\ \quad superfluid pair breaking & \cite{Voskresenskii87a}, \\ & \cite{Schaab95b} \\ \hline Photosphere: & \cite{Potekhin96c} \\ \noalign{\smallskip}\hline \end{tabular} \end{table*} \subsection{Equation of state} \label{sec:inp.eos} For the outer and inner crust we adopt the equations of state of Haensel \& Pichon (1994)\nocite{Haensel94a} and Negele \& Vautherin (1973)\nocite{Negele73}. The transition density between the ionic crust and the core of a neutron star is taken to be $\rho_{{\rm tr}}=1.7\times 10^{14}\mathrm{\,g\,cm}^{-3}$ (\cite{Pethick95}). For the present study, we choose two models of high density matter. The first, non-relativistic model (UV$_{14}$+UVII) is obtained by solving the Schr\"odinger equation by means of a variational approach (\cite{Wiringa88}). The other model (RHF8) uses relativistic Br\"uckner-Hartree-Fock results up to 2-3 times normal nuclear density. Hyperons are included within the relativistic Hartree-Fock approach above this density (\cite{Huber97a}). One important difference between these equations of state is that the non-relativistic model treat neutron star matter as being composed of neutrons and protons only (which are in $\beta$-equilibrium with leptons), whereas the relativistic model accounts for all hyperon states that become populated in the cores of neutron stars. \subsection{Neutrino emissivity} \label{sec:inp.emis} The neutrino emission processes can be divided into slow and enhanced processes depending on whether one or two baryons participate in the reaction. Due to the rather different phase spaces associated with both kind of processes the emission rates differ by several orders of magnitude. If one neglects exotic states in neutron star matter (like quark-gluon plasma and meson condensates) the only enhanced neutrino emission processes are the direct nucleon and hyperon Urca processes \begin{align} \mathrm{n} &\rightarrow \mathrm{p} + \lambda^- + \bar{\nu}_\lambda \\ B_1 &\rightarrow B_2 + \lambda^- + \bar{\nu}_\lambda , \end{align} respectively, where $B_{1,2}=\mathrm{n}$, p, $\Sigma^{\pm,0}$, $\Lambda$, $\Xi^{0,-}$ denotes the baryons and $\lambda_{1,2}=\mathrm{e}^-$, $\mu^-$ the leptons. Because of the $\beta$-equilibrium the inverse reaction (with $\bar\nu$ replaced by $\nu$) occurs at the same rate as the direct one. The emissivities of these processes were computed by Prakash et al. (1992)\nocite{Prakash92}. Simultaneous conservation of energy and momentum requires that the triangle inequality $p_{B_1}^{\rm F} < p_{B_2}^{\rm F} + p_{\rm e}^{\rm F}$ and the two inequalities obtained by cyclic permutation are fulfilled for the Fermi momenta $p_i^{\rm F}$. If the inequalities are not fulfilled the process is extremely unlikely to occur and the corresponding emissivity can be neglected. The availability of the various fast processes thus depends on the partial concentrations of each baryon species. The proton fraction, which determines the possibility of the direct nucleon Urca process, depends crucially on the symmetry energy, which is poorly known for high density matter. As an example of an equation of state which yields slow cooling we study the non-relativistic equation of state UV$_{14}$+UVII. Due to the non-monotonic behaviour of the symmetry energy at high densities, the triangle inequality is not fulfilled for this equation of state. The relativistic equation of state RHF8 allows for both the direct nucleon and the direct hyperon Urca processes for star masses $M>1.09\,M_\odot$ and $M>1.22\,M_\odot$, respectively. During the transition to the superfluid state of neutrons and protons the superfluid pair breaking and formation processes become important (\cite{Flowers76,Voskresenskii87a}). We refer to Schaab et al. (1997)\nocite{Schaab95b} for a detailed description of these processes and their impact on the cooling of neutron stars. \subsection{Superfluidity} \label{sec:sf} The pairing gaps are sensitive to the underlying microscopic model for the nucleon-nucleon interaction. Since these models are rather uncertain, especially at high densities, the value of the maximum gap and the density range where pairing can occur is model dependent. The disagreement arising from the different interaction models is enlarged by more subtle issues of the medium polarisation (\cite{Wambach91a,Schulze96a}) or the inclusion of relativity (\cite{Elgaroy96c}). \begin{table*} \centering \caption[]{Gap energies of superfluid states in neutron star matter. The density ranges are calculated for the RHF8 equation of state. \label{tab:sf}} \begin{tabular}{lccl} \hline Pairing & $\Delta_{\rm max}$ [MeV] & Density range [fm$^{-3}$] & References \\ \hline neutron $^1\mathrm{S}_0$ & 1.01 & $<\,0.16$ & \cite{Schulze96a} \\ proton $^1\mathrm{S}_0$ & 0.92 & 0.10--0.31 & \cite{Elgaroy96a}\\ neutron $^3\mathrm{P}_2$ & 1.45 & $>\,0.06$ & \cite{Baldo98a}, CD-Bonn potential \\ lambda $^1\mathrm{S}_0$ & 0.24 & 0.44--0.63 & \cite{Balberg97b} \\ \hline \end{tabular} \end{table*} At densities of the order $2\times 10^{14}\mathrm{\,g\,cm}^{-3}$ the $S$ state interaction becomes repulsive for neutrons, and the $^1S_0$ neutron gap closes. The attractive $^3P_2-^3F_2$ state interaction leads to the pairing in this channel at about the same densities. For the $P$ state pairing we use the recent calculation of Baldo et al. (1998)\nocite{Baldo98a} which is based on the modern nucleon--nucleon CD-Bonn potential (\cite{Machleidt96a}), which provides accurate fits to the actual nucleon-nucleon scattering data below 350~MeV. Field theoretical descriptions of the heavy-baryon interaction based on the one-boson-exchange model indicate that the $\Lambda-\Lambda$ interaction has an attractive component. From the so-called doubly-strange hypernuclei, where two hyperons are bound in a single nucleus (\cite{Imai92a}), one can verify this field theoretical result, since, for instance, the separation energy of the two $\Lambda$ hyperons, $B_{\Lambda\Lambda}$, exceeds the separation energy of a single $\Lambda$ from the same nucleus, $B_{\Lambda}$, by more than a factor of two. As expected according to the quark model, the binding energy of hyperons, $\Delta B_{\Lambda\Lambda}\!\equiv\!B_{\Lambda\Lambda}-2B_\Lambda\approx4-5\;$ MeV, is somewhat less than the corresponding binding energy of nucleons, $\Delta B_{NN}\approx6-7\;$MeV. Due to these similarities, $\Lambda$ hyperon pairing is expected in high density matter in analogy with the case of nucleons. Here, we use the $\Lambda$ hyperon $^1\mathrm{S}_0$\ gaps estimated by Balberg \& Barnea (1998)\nocite{Balberg97b}. The impact of $\Lambda$ pairing on the thermal evolution of neutron stars has recently been investigated in Schaab et al. (1998a)\nocite{Schaab98b}. After the onset of the superfluid state of the respective constituents of stellar matter the neutrino emissivities, the thermal conductivity, and the heat capacity are suppressed by factors which behave like $\exp(-aT_{\rm c}/T)$ for $T\ll T_{\rm c}$, where $T_{\rm c}$ denotes the critical temperature and $a$ is some constant of order unity (\cite{Maxwell79,Levenfish94a,Gnedin95a,Yakovlev95a}). We refer to Levenfish \& Yakovlev (1994)\nocite{Levenfish94a} for fitting formulas of ${\mathcal R}_{\rm sf}$, that are valid over the whole temperature range $T<T_{\rm c}$ for both types of pairing state. The suppression factor of the $^3\mathrm{P}_2$\ pairing behaves exponentially only if the neutrons pair in the nodeless state with $m=0$. In the case of $m=2$ this factor behaves polynomially (\cite{Anderson61,Muzikar80,Levenfish94a}). The effects of $m=2$ pairing on the cooling behaviour is discussed in Schaab et al. (1998b)\nocite{Schaab98a} where considerable modifications were found. Due to the uncertainty in of the ground state quantum numbers of the neutron condensate we shall adopt here the more conventional case of the $m= 0$ nodeless pairing. \subsection{Photosphere} \label{sec:inp.photo} While the heat conduction in the interior of the star is treated dynamically in our simulations the actual surface temperatures are obtained by attaching an outer envelope with $\rho \le \rho_{\rm m}$ at the boundary to the 'core'. The validity of this approximation has been discussed in Gudmundsson et al. (1983)\nocite{Gudmundson83}. To obtain the surface temperatures we use the envelope calculations by Potekhin et al. (1997)\nocite{Potekhin96c} whose non-magnetic photospheres provide a smaller gradient for low effective surface temperatures $T_{\rm eff}\lesssim 3\times 10^5{\rm\,K}$ than the earlier calculations of Gudmundson et al. (1983)\nocite{Gudmundson83} and Van Riper (1988)\nocite{VanRiper88}, mainly due to a refined treatment of the electron conductivity in the non-degenerate region. This yields a slower cooling in the photon cooling era for stars older than $\gtrsim10^6{\rm\,yr}$. Even a rather small amount of accreted matter of say $\Delta M\sim10^{-16}M_\odot$, substantially reduces the thermal insulation of the photosphere. The effect is highest for $\Delta M\sim10^{-7}M_\odot$, where the accreted hydrogen is partly burned into helium and carbon. Additionally accreted matter will be converted into iron, leaving the thermal insulation of the star practically unchanged (\cite{Potekhin96c}). The effect of accreted envelopes on cooling of neutron stars are discussed in Sect. \ref{sec:res.accr}. \subsection{Comparison of the efficiency of the heating mechanism} \label{sec:inp.comp} In Fig. \ref{fig:lum} the heating rates \begin{equation} H = \int 4\pi r^2 {\rm e}^{\Lambda+2\phi} h {\rm d} r \end{equation} discussed in Sect. \ref{sec:vortex} and \ref{sec:nonequilibrium} are compared with the total photon and neutrino luminosities $L_\gamma$ and $L_\nu$, respectively, where an isothermal configuration, i.e. $T{\rm e}^\phi =\mathrm{const.}$, is assumed. Since the heating rates depend not only on the temperature but also on the time, we use the standard thermal evolution scenario without heating to link the star's temperature to its age. \begin{figure*} \centering\resizebox{0.8\hsize}{!}{\rotatebox{-90}{\includegraphics{h1252.f2}}} \caption{Heating rate $H$ as functions of internal temperature for A: EB-pinning, B: PVB-pinning, C: corotating crust, D: corotating core, E: crust cracking, F: chemical heating of the crust. The left panel corresponds to $K=10^{-15}$~s, the right one to $K=10^{-22}$~s. The photon luminosity, as well as the neutrino luminosity of a standard cooling and an enhanced cooling model are also shown.} \label{fig:lum} \end{figure*} We use two values of the rotational parameter $K$ (see Eq. \eqref{eq:rot.evol}). About the first value, $K=10^{-15}$~s (left panel in Fig. \ref{fig:lum}), the observed values of most of the pulsars scatter, particularly of the five pulsars in class A. However there are four pulsars (see Table \ref{tab:observation}) which are rather old, $\tau >6\times 10^7$~yr, and rotate very fast, $P<6$~ms. These pulsars are supposed to be \emph{recycled} by accretion of matter from a companion star (see, e.g., \cite{Lorimer96a}). Since their rotational parameter $K\sim10^{-22}$~s is quite different from the one of all the other pulsars, we shall study them separately (see Sect. \ref{sec:res.recycled} and right panel of Fig. \ref{fig:lum}). For the value $K=10^{-15}$~s the rate of only three heating mechanism, EB- and PVB-pinning and corotating core, exceed both, the photon and the neutrino luminosities. We expect (and obtain) therefore only for these three heating mechanism a considerable effect (see Sect. \ref{sec:res}). All these heating mechanism affect both the late and the early cooling. This is in contrast to the case, where $K=10^{-22}$~s is employed. Since the time derivative of the angular velocity $\dot\Omega$ is small in the latter case, only the late cooling is affected. Besides the three heating mechanisms quoted above, the other heating mechanisms result in a heating rate which is larger than the photon luminosity, except the heating from a corotating crust. The rate of chemical heating of the core depends also on the deviation from chemical equilibrium and can be compared with the other processes only in detailed cooling simulations. As we will see in Sect. \ref{sec:res.recycled}, it considerably affects the late cooling for $K=10^{-22}$~s. \section{Observed data} \label{sec:observations} Only a few of the known pulsars have been detected by the X-ray observatories Einstein, EXOSAT, ROSAT, and ASCA during the last two decades. We use a sample of 27 pulsars to compare the theoretical cooling curves with the observed data (see Table \ref{tab:observation}). Both the timing characteristics (i.e. $\Omega$ and $\dot\Omega$) and at least an upper limit on the effective surface temperature $T_{\rm eff}^\infty$ as measured at infinity are known for these pulsars. Table \ref{tab:observation} summarises these data. The effective surface temperatures are specified together with their $2\sigma$ error range. \begin{table*} \centering \caption[]{Sample of observed data \label{tab:observation}} \small \renewcommand{\baselinestretch}{1.0} \begin{tabular}{lccccccll} \hline Pulsar & $P$ & $\dot P$ & $\log(\tau)$ & $\log(K)$ & $\log(T_{\rm eff}^\infty)$ & Category & Reference \\ & [ms] & [$10^{-15}{\rm\,ss}^{-1}$] & [yr] & [s] & [K] & & \\ \hline 0531+21 & 33.40 & 420.96 & 2.97\dag & -13.9 & $6.18^{+0.19}_{-0.06}$ & B & 1 \\ (Crab) &&&&&&& \\ 1509-58 & 150.23 & 1540.19 & 3.19 & -12.6 & $6.11\pm 0.10$ & B & 2 \\ 0540-69 & 50.37 & 479.06 & 3.22 & -13.6 & $6.77^{+0.03}_{-0.04}$ & B & 3,4 \\ 0002+62 & 241.81 & & $\sim 4$\dag & $\sim -13$ & $6.20^{+0.07}_{-0.27}$ & A,bb & 5 \\ 0833-45 & 89.29 & 124.68 & $4.3\pm 0.3$\dag & -14.0 & $6.24\pm0.03$ & A,bb & 6 \\ (Vela) &&&&& $5.88\pm 0.09$ & A,mH & 7 \\ 1706-44 & 102.45 & 93.04 & 4.24 & -14.0 & $6.03^{+0.06}_{-0.08}$ & B & 8 \\ 1823-13 & 101.45 & 74.95 & 4.33 & -14.1 & $6.01\pm 0.02$ & C & 9 \\ 2334+61 & 495.24 & 191.91 & 4.61 & -13.0 & $5.92^{+0.15}_{-0.09}$ & C & 10 \\ 1916+14 & 1181 & 211.8 & 4.95 & -12.6 & $5.93$ & D & 11 \\ 1951+32 & 39.53 & 5.85 & 5.03 & -15.7 & $6.14^{+0.03}_{-0.05}$ & B & 12 \\ 0656+14 & 384.87 & 55.03 & 5.05 & -13.7 & $5.98\pm0.05$ & A,bb & 13 \\ &&&&& $5.72^{+0.06}_{-0.05}$ & A,mH & 14 \\ 0740-28 & 167 & 16.8 & 5.20 & -14.6 & $5.93$ & D & 11 \\ 1822-09 & 769 & 52.39 & 5.37 & -13.4 & $5.78$ & D & 11 \\ 0114+58 & 101 & 5.84 & 5.44 & -15.2 & $5.98\pm 0.03$ & C & 11 \\ 1259-63 & 47.76 & 2.27 & 5.52 & -16.0 & $5.88$ & C & 15 \\ 0630+18 & 237.09 & 10.97 & 5.53 & -14.6 & $5.76^{+0.04}_{-0.08}$ & A,bb & 16 \\ (Geminga) &&&&& $5.42^{+0.12}_{-0.04}$ & A,mH & 17 \\ 1055-52 & 197.10 & 5.83 & 5.73 & -14.9 & $5.90^{+0.06}_{-0.12}$ & A,bb & 18 \\ 0355+54 & 156.38 & 4.39 & 5.75 & -15.2 & $5.98 \pm 0.04$ & C & 19 \\ 0538+28 & 143.15 & 3.66 & 5.79 & -15.3 & $5.83$ & C & 20 \\ 1929+10 & 226.51 & 1.16 & 6.49 & -15.6 & $5.52$ & B & 21 \\ 1642-03 & 388 & 1.77 & 6.54 & -15.2 & $6.01\pm 0.03$ & C & 11 \\ 0950+08 & 253.06 & 0.23 & 7.24 & -16.3 & $4.93^{+0.07}_{-0.05}$ & B & 22 \\ 0031-07 & 943 & 0.40 & 7.56 & -15.4 & $5.57$ & D & 11 \\ 0751+18 & 3.47 & $7.9\times 10^{-4}$ & 7.83 & -20.6 & $5.66$ & C & 23 \\ 0218+42 & 2.32 & $8.0\times 10^{-5}$ & 8.66 & -21.7 & $5.78$ & C & 24 \\ 1957+20 & 1.60 & $1.7\times 10^{-5}$ & 9.18 & -22.6 & $5.53$ & C & 25,26 \\ 0437-47 & 5.75 & $3.8\times 10^{-5}$ & 9.20 & -21.7 & $5.94\pm 0.03$ & B & 27 \\ \hline \end{tabular} \comments{The entries are: rotation period $P$, spin-down age $\tau=P/2\dot P$, $K=P\dot P$ (see Eq. \eqref{eq:rot.evol}), effective surface temperature as measured at infinity $T_{\rm eff}^\infty$. The four categories A to D are explained in the text. bb and mH refer to blackbody and magnetic hydrogen atmosphere fits, respectively. \dag: estimated true age instead of spin-down age (see text). References: 1: \cite{Becker95a}, 2: \cite{Seward83a}, 3: \cite{Finley93a}, 4: \cite{Boyd95a}, 5: \cite{Hailey95a}, Fig. 2 with lower limit on $N_{\rm H}$, 6: \cite{Oegelman95a}, Table III, 7: \cite{Page96a}, Fig. 1, 8: \cite{Becker95b}, 9: \cite{Finley93b}, 10: \cite{Becker96a}, 11: \cite{Slane95a}, 12: \cite{SafiHarb95b}, 13: \cite{Possenti96a}, 14: \cite{Anderson93}, 15: \cite{Becker96c}, 16: \cite{Halpern97a}, PSPC+SIS0, 17: \cite{Meyer94}, Fig. 2a with $B_{12}=1.18$, 18: \cite{Greiveldinger96a}, Table 2, 19: \cite{Slane94a}, 20: \cite{Sun96a}, 21: \cite{Yancopoulos94a}, 22: \cite{Manning94a}, 23: \cite{Becker96d}, 24: \cite{Verbunt96a}, 25: \cite{Kulkarni92a}, 26: \cite{Fruchter92a}, 27: \cite{Zavlin97a}.} \end{table*} The ages of all pulsars except PSRs 0531+21 (Crab), 0833-45 (Vela) and 0002+62 are estimated via their spin-down age, $\tau=P/2\dot P$. This relation requires that both the moment of inertia and the magnetic surface field are constant, and that the braking index $n$ is equal to its canonical value of 3, which corresponds to a spin-down by emission of pure magnetic dipole radiation (see Sect.\ \ref{sec:spin.normal}). Of course, the spin-down age is a crude approximation to the true age of a neutron star, which can deviate from this value by a factor as large as $\sim 3$, as discussed before. The situation is different for the three pulsars quoted above, where ages are known from different sources, that is: the age of the Crab pulsar is known from the chronicles, the age of Vela was recently determined via the proper motion of the pulsar with respect to the supernova remnant by Aschenbach et al. (1995)\nocite{Aschenbach95a}, and the approximate age of PSR 0002+62 is given by an estimate of the age of the related supernova remnant G 117.7+06 (\cite{Hailey95a}). The information obtained from the X-ray observations is not always sufficient to extract the effective surface temperature of the corresponding neutron star. The sample is therefore divided into four categories labelled A through D (\cite{Oegelman95a,Schaab95a}): \begin{description} \item[Category D:] The four pulsars have not been detected in the soft X-ray range so far. By considering the instrumental sensitivity an upper limit for the surface temperature could be set. These pulsars are marked with white triangles in the figures below. \item[Category C:] The detections of ten pulsars contain too few photons for spectral fits. The surface temperatures were obtained by using the totally detected photon flux. These pulsars are marked with black triangles. \item[Category B:] The spectra of eight pulsars, including the Crab pulsar 0531+21, can only be fitted by a power--law--type spectrum or by a blackbody spectrum with very high effective temperatures and effective areas much smaller than a neutron star surface. Presumably, their X-ray emission is dominated by magnetospheric emission. Therefore, the temperatures, determined from the spectral fits, are probably higher than the actual surface temperatures. Pulsars of this type are marked with arrows. \item[Category A:] Finally, there are five pulsars, 0833-45 (Vela), 0656+14, 0002+62, 0630+18 (Geminga), and 1055-52, which allow two--component spectral fits. The softer blackbody component is believed to correspond to the actual surface emission of the neutron star, while the harder blackbody (or power--law) component may be due to magnetospheric emission. These pulsars are marked with errorbars. \end{description} The obtained effective surface temperature of pulsars of category A depend crucially on whether or not a hydrogen atmosphere is assumed. PSR 0833-45 (Vela) is a specific example (cf. Table \ref{tab:observation}). Generally a spectral fit with a magnetic hydrogen atmosphere yields a substantially lower effective surface temperature than a blackbody fit. These models of hydrogen atmospheres however do not yet take into account the presence of neutral atoms (\cite{Potekhin96c}). Their effect should yield a spectrum that is more similar to the blackbody spectrum, as indicated by the preliminary estimates by Shibanov et al. (1993)\nocite{Shibanov93a}. The atmospheric composition of a specific pulsar could be determined by considering multiwavelength observations in the near future (\cite{Pavlov96a}). We shall use the effective surface temperatures obtained by fitting the spectra with both the blackbody and the magnetic hydrogen model atmosphere for the three pulsars of category A, for which both fits were accomplished. \section{Results and discussion} \label{sec:res} \subsection{Standard cooling} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{h1252.f3}} \caption[]{Standard cooling with and without internal heating. Parameters are: UV$_{14}$+UVII\ equation of state, $M=1.4M_\odot$, and $K=10^{-15}$~s. See Table \ref{tab:observation} for the observed data. \label{fig:cool1}} \end{figure} Figure \ref{fig:cool1} shows the surface temperature as measured by a distant observer as a function of the star's age for the slowly cooling neutron star models with and without internal heating. These models are based on the non-relativistic equation of state UV$_{14}$+UVII\ which account only for chemically equilibrated nucleons and leptons (cf. Sect. \ref{sec:inp.eos}). We choose the canonical value $M=1.4M_\odot$ for the gravitational mass of a neutron star. Note that the relatively low surface temperature is caused by the inclusion of the superfluid pair breaking and formation process (\cite{Flowers76b,Voskresenskii87a,Schaab95b}), which was not taken into account in the earlier investigations (e.g. \cite{Umeda94}; \cite{Page95}; \cite{Schaab95a}). The standard cooling model without internal heating is consistent with most of the observations. A discrepancy occurs however for the effective surface temperature of the Vela pulsar 0833-45 derived for a blackbody fit, which tends to lie considerably above all cooling tracks. The effective radius obtained for the blackbody fit, i.e. $R_{\rm eff}\approx 6{\rm\, km}$, is considerably smaller than the canonical value of $\sim 10 {\rm km}$, generally obtained for standard equations of state. Therefore, the existence of a magnetic hydrogen (or helium) atmosphere is more reliable and in agreement with the standard cooling scenario without internal heating. The effect of heating of the crust due to thermal creep of pinned vortices is shown by the dotted and by the short dashed curves in Fig. \ref{fig:cool1} for the EB- and PVB-pinning model, respectively (see Sect. \ref{sec:pinning}). The heating rate depends on $\dot\Omega$ (see Eq. [\ref{eq:pinn.heatrate}]) and thus on the inverse of the magnetic dipole moment. The employed value $K=10^{-15}$\,s (see Eq. [\ref{eq:rot.evol}]) is in the range $10^{-16}$--$10^{-13}$\,s about which the observed values tend to scatter, with the exception of the four millisecond pulsars which will be addressed in Sect. \ref{sec:res.recycled}. The effective surface temperature is increased in both models in a similar way, though the effect of the EB-pinning model is larger because of the larger pinning energy in the inner crust. The heating of the crust can easily be recognised by the increase of the thermal diffusion time, i.e. by the amount of time needed for the cooling wave to reach the surface (\cite{Lattimer94a}). After this time the surface temperature drops significantly. It is increased by a factor of $\sim 50$ ($\sim 10$) in the case of EB- (PVB-)pinning. This characteristic feature will become very interesting if the thermal spectrum of a young pulsar $\tau\lesssim 100{\rm\, yr}$ should be detected. For $\tau\gtrsim 10^6{\rm\, yr}$ the slope of the cooling track is increased by the heating of the crust. The observed upper limit of the effective surface temperature of PSR 0950+08 seems to rule out the EB-pinning model. Although consideration of a magnetic photosphere would decrease the slope again, compensation of the heating effect would require an unphysically strong magnetic field (\cite{VanRiper88}). For middle aged pulsars, $10^3{\rm\, yr}<\tau<10^6{\rm\, yr}$, heating due the PVB-pinning model has only a small effect compared to the other uncertainties of the input parameters. The other three heating processes possible in the crust, driven by drag forces, crust cracking, and chemical heating have no visible effect on the cooling, since the heating rates are too small in comparison with the neutrino and photon emission rates (for the value of $K$ used here). In the case of crust cracking our results are in contrast to the ones reported by Cheng et al. (1992)\nocite{Cheng92a}, where a significant effect especially during the late cooling epoch has been obtained. This deviation can be traced back to different expressions used for the heating rates (see Sect. \ref{sec:noneq.cracking}) Electron-vortex scattering in the interior of the star affects the stellar cooling at times in the range $10^1<\tau<10^6{\rm\, yr}$ (see Fig. \ref{fig:cool1}). During this period the effect is comparable to the effect of EB-pinning in the crust. Heating due to vortex decay at the crust-core boundary yields an increase of the surface temperature only for rather old pulsars ($\tau>10^7{\rm\, yr}$). It can not be constrained by any present pulsar observation, however. \subsection{Enhanced cooling} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{h1252.f4}} \caption[]{Enhanced cooling of neutron star models constructed for the RHF8 equation of state. Other parameters are as in Fig. \ref{fig:cool1}. \label{fig:cool4}} \end{figure} Figure \ref{fig:cool4} shows the cooling behaviour of enhanced cooling models which are based on the RHF8 equation of state. The surface temperature drops by a factor of $\sim 6$ when the cooling wave has reached the surface. The fast direct Urca processes are only suppressed below the critical temperature of neutron and lambda pairing, $T_{\rm c}\sim2.0\times 10^9$~K and $T_{\rm c}\sim 1.6\times 10^9$~K, respectively. The effect of heating on the cooling behaviour is similar as in the case of standard cooling. The EB-pinning model shows again the largest effect on the surface temperature. With the only exception of PSR 0630+18 (Geminga), the effective surface temperatures of the pulsars of category A are clearly not consistent with the enhanced cooling models, no matter whether one includes heating processes or not. This discrepancy can be reduced either by enhancing the gap energies (\cite{Page95,Schaab95a}) or by assuming an accreted envelop. \subsection{Accretion} \label{sec:res.accr} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{h1252.f5}} \caption[]{Standard cooling with fully accreted envelope. Other parameters are as in Fig. \ref{fig:cool1}. \label{fig:cool5}} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{h1252.f6}} \caption[]{Enhanced cooling with fully accreted envelope. Other parameters are as in Fig. \ref{fig:cool4}. \label{fig:cool7}} \end{figure} In Figs. \ref{fig:cool5} and \ref{fig:cool7}, we study the effect of a fully accreted envelope, which means that the accreted mass exceeds $\sim 10^{-7}M_\odot$ (see \cite{Potekhin96c} and Sect. \ref{sec:inp.photo}). Due to the envelope consisting of light atoms such as hydrogen, helium and carbon, the temperature gradient of the photosphere is substantially reduced. This yields a high effective surface temperature in the neutrino cooling era, during which the energy loss rate depends on the internal temperature; whereas it yields small surface temperatures in the photon cooling era where the loss rate depends on the surface temperature itself. As it was already noted, the enhanced cooling models are now in better agreement with the data of observation class A (see Fig. \ref{fig:cool7}). This is especially true for the surface temperatures obtained within the magnetic hydrogen atmosphere model. The enhanced cooling model with heating due PVB-pinning fits almost perfectly these surface temperatures. \subsection{Millisecond pulsars} \label{sec:res.recycled} So far we have investigated cooling scenarios associated with a large rotational parameter, i.e. $K=10^{-15}$~s (see Eq. [\ref{eq:rot.evol}]). This value is consistent with the bulk fraction of observed pulsars (cf. Table \ref{tab:observation}). The four oldest pulsars, PSR 0751+18, 0218+42, 1957+20, and 0437-47, however have much smaller values around $K\sim 10^{-22}$~s. As it was shown in Sect. \ref{sec:inp.comp}, this causes considerably different heating rates. In particular, the mechanism of chemical heating and crust cracking then becomes efficient enough to noticeably influence the thermal evolution. In Fig. \ref{fig:cool6} we show the results for the various heating mechanism in comparison with models without heating. Since the late cooling behaviour (for $\tau\gtrsim 10^7$~yr) does not depend on the cooling scenario, we consider here only the standard cooling scenario. The model with chemical heating of the core shows a sudden rise of the surface temperature at $\tau\sim 10^6$~yr. This is caused by the release of chemical energy through the $\beta$-decay of neutrons. This process is suppressed by superfluidity if the deviation from chemical equilibrium is smaller than the gap energy (s. Sect. \ref{sec:chem} and \cite{Reisenegger96a}). \begin{figure} \resizebox{\hsize}{!}{\includegraphics{h1252.f7}} \caption[]{Same as Fig. \ref{fig:cool1} but with various heating mechanism and $K=10^{-22}$~s. The curves correspond to the models without internal heating (labeld with A), with EB-pinning (B), PVB-pinning (C), crust cracking (D), corotating core (E), chemical heating of the crust (F), and chemical heating of the core (G), respectively. \label{fig:cool6}} \end{figure} All cooling tracks in Fig. \ref{fig:cool6} are consistent with the upper temperature limits of the three oldest millisecond pulsars. Only the observation of PSR 0751+18 seems to exclude the EB-pinning model. The situation changes however if one assumes that the actual temperatures are not too far from the upper limits. Then only the PVB- and EB-pinning models, as well as the chemical heating of the superfluid core yield sufficiently high surface temperature. A future temperature determination, which will set a firm lower limit on the temperature for one of these pulsars could therefore rule out at least the corotation and the crust cracking models as the only heat sources in millisecond pulsars. \subsection{Two dimensional simulations} In this section we relax the simplification that the heating rates are angle independent and extend our one-dimensional calculations of the previous sections to the two-dimensional case. The averaging of the local heating rates over spherical shells assumes that the heat conductivity is very high in transverse direction. This is indeed a fairly good approximation, as we shall see next by comparing the results of one dimensional calculation with those of fully two-dimensional ones. The two-dimensional code was developed recently by Schaab \& Weigel (1998)\nocite{Schaab98c} (s.a. \cite{Schaab97a}). Fig. \ref{fig:cool8} shows the obtained cooling tracks for standard cooling computed for the UV$_{14}$+UVII --model with inclusion of the EB-pinning model. Heating leads to a large temperature difference (up to 50\,\%) between the pole ($\theta=0$) and the equator ($\theta=\pi/2$) for a pulsar whose age lies in the range $50\lesssim\tau\lesssim10^4$~yr. Moreover the temperature drop at the equator is delayed by the heating with respect to the model without heating and with respect to the drop of the polar temperature. This is the result of the $h\propto\sin\,\theta$ dependence for the vortex creep model. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{h1252.f8}} \caption[]{Comparison of the one and two dimensional cooling simulations for the UV$_{14}$+UVII --model with inclusion of EB-pinning. \label{fig:cool8}} \end{figure} The surface temperature obtained with the one dimensional code can be compared with the effective temperature which is defined by (\cite{Page95c}) \begin{equation} T_{\rm eff} = \left(\int_0^1 {\rm d} s ~ 2 s ~ \left( {\rm e}^\nu T(\theta)\right)^4 \right)^{1/4}, \end{equation} where $s=\sin\,\delta$. The emission angle $\delta$ is related to the colatitude $\theta$ by the equation \begin{equation} \label{eq:geodes} \theta(\delta) = \int_0^{M/R}{\rm d} u \frac{\sin\,\delta} {\left(\left(1-\frac{2M}{R}\right)\left(\frac{M}{R}\right)^2 -(1-2u)u^2\sin^2\,\delta\right)^{1/2}}, \end{equation} where $M$ and $R$ are the neutron star's mass and radius, respectively. In flat spacetime, equation \eqref{eq:geodes} simplifies to $\delta =\theta$. The effective temperature $T_{\rm eff}$ and the surface temperature obtained with the one dimensional code are almost identical (see Fig. \ref{fig:cool8}). This means that the luminosity obtained with both codes are also identical. The results of the one dimensional codes are thus acceptable for a comparison with the observed data of categories B--D, where the upper bounds on the surface temperature are derived from the respective luminosity. If one uses spectral information as for the pulsars in category A, the observations should be compared with the spectra derived from the surface temperature distribution $T_{\rm s}(\theta)$ in the framework of two dimensional simulations. Nevertheless, we expect that the conclusions remain unchanged if we use the effective temperature also for a comparison with the observed data of category A. \section{Conclusions} \label{sec:disc} In this work, we have carried out a comparative analysis of the impact of different competing heating processes on the cooling behaviour of neutron stars. In general we find that the internal heating yields significantly enhanced surface temperatures of middle aged and old pulsars, as one would expect. However, the effectiveness of these heating processes varies significantly from one process to another, and depends sensitively on the value of the rotational parameter $K$. We studied models with two different rotational parameters. The first value was chosen to amount $K\sim 10^{-15}$~s, which is supported by pulsars observations. For this $K$ value, the heating due to thermal creep of pinned vortices and motion of proton vortices in the interior of the star yield considerable enhancements of the surface temperatures of middle aged pulsars with respect to the models which ignore heating. This leads to closer agreement with the observed data in the case of enhanced cooling, which is even more improved for the case of the fully accreted envelopes (s.a. \cite{Page96b}). The observed upper temperature limit for PSR 0950+08 seems to rule out the strong pinning models, whereas the weaker pinning models are consistent with the observations (the pinning models differ by the value of the pinning energy in the inner crust). The other heating processes -- from chemical heating to dissipative motion of neutron vortices in the crust to crust cracking -- have no observable effect on the cooling. Our result for heating due to crust cracking is in contrast to the one obtained by Cheng et al. (1992)\nocite{Cheng92a}, who find a large effect especially on the late cooling stages. The difference arises from different expressions adopted for the heating rates. The four oldest pulsars of our sample rotate very fast with periods of a few milliseconds. Hence the rotational parameter for millisecond pulsars is smaller than that for the bulk fraction of the population by 7 orders of magnitude ($K=10^{-22}$~s). For this value only the late cooling stages are affected by heating. We find that the heating due to thermal creep of pinned vortices and chemical heating of the core has the largest impact on the surface temperatures of millisecond pulsars. Again, the strong pinning models leads to surface temperatures which are larger than one of the upper limits deduced for the millisecond pulsars. As far as we know, the published cooling simulations of neutron stars which account for internal heating were performed in one spatial dimension. Such a treatment is only possible if the angle dependent heating rates are averaged over spherical shells. By comparing the outcome of one-dimensional simulations with fully two-dimensional simulations, based on a recently developed numerical code (\cite{Schaab98c}), we could confirm the validity of this procedure. The knowledge of the input parameters for cooling simulations such as the composition, the superfluid gap energies, and the possible existence of an accreted envelope, etc. is far form being complete. The number of parameters involved in the cooling simulations increases when the internal heating processes are taken into account. In principal, one could expect to derive some of the parameters associated with these phenomena by means of comparing the theoretical cooling models with the body of observed data. This attempt needs to be supplemented with the analysis of terrestical experiments, other astrophysical observations, as well as future theoretical studies. Nevertheless, the observed data already constrain the microphysical input, as, for instance, for the scenaria of heating via vortex creep. Tables containing detailed references to the ingredients used in the simulations, the observational data, and the resulting cooling tracks can be found on the Web: http://www.physik.uni-muenchen.de\hspace{0pt}/sektion\hspace{0pt}/suessmann\hspace{0pt}/astro\hspace{0pt}/cool. \section*{Acknowledgments} We would like to thank H.-T. Janka for helpful discussions and the anonymous referee for many valuable suggestions. Ch.~S. gratefully acknowledges the Bavarian State for a fellowship. A.~S. has been supported through a research grant from the Max Kade Foundation.
\section*{III. Exact solutions} Remarkably, for the equation under study it is possible to give a complete account of solutions with separated variables. They have the form (\ref{2.1}) and the separation equations for the functions $\vp_\mu,\ (\mu=0,1,2,3)$ read as (\ref{2.l}), where the coefficients $F_{ai},\ a,i=1,2,3$ are the entries of the corresponding St\"ackel matrices (\ref{2.s}), functions $T_a,\ a=1,2,3$ are listed in (\ref{5a}) and the functions $T_0, F_{a0}, a=1,2,3$ given in (\ref{4b}). The separation equation for the function $\vp_0(t)$ is easily integrated. The separation equations for the functions $\vp_i(\om_i),\ (i=1,2,3)$ are similar to those arising from separation of variables in the Helmholtz equation $(\Delta_3+\om^2)\Psi=0$. The solutions of these equations are well known (see, \cite{mil77,mor53} and the references therein). Below we adduce solutions of FPE (\ref{1.1}) for each class of functions $\vec z = \vec z (\vec \om)$ given in (\ref{1.2}). \begin{enumerate} \item{Cartesian coordinates \[ u(t,\vec \om)=\exp\left\{\sum\limits_{i=1}^{3}\, \left(\lambda_i {c_i^{-2} \over 2 l_i}e^{-2l_it}-l_i t\right)\right\}\exp(i(\al\om_1+\be\om_2+\ga\om_3)) \] and $ \la_1=-\al^2, \la_2=-\be^2, \la_3=-\ga^2.$} \item {Cylindrical coordinates \begin{eqnarray*} &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}+\lambda_3 {c_3^{-2}\over 2 l_3}e^{-2l_3t}- (2l_1+l_3)t\right\}\times\\ &&\quad J_n(\al e^{\om_1})\exp(i(n\om_2+\ga\om_3)), \end{eqnarray*} where $J_n$ is the Bessel function \cite{wat,erd}, and $\la_1=-\al^2, \la_2=-n^2, \la_3=-\ga^2.$} \item {Parabolic cylindrical coordinates \begin{eqnarray*} &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}+\lambda_3 {c_3^{-2}\over 2 l_3}e^{-2l_3t}- (2l_1+l_3)t\right\}\times\\ &&\quad D_{i\mu-1/2}(\pm\sigma\om_1)\, D_{-i\mu-1/2}(\pm\sigma\om_2)\, e^{i\om_3\ga}, \end{eqnarray*} where $\sigma=e^{i\pi/4}(2\al)^{1/2}, D_\nu$ is the parabolic cylinder function \cite{buc,erd} and $ \la_1=-\al^2, \la_2=-2\al\mu, \la_3=-\ga^2.$} \item{For the case of elliptic cylindrical coordinates we have two types of solutions \begin{eqnarray*} &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}+\lambda_3 {c_3^{-2}\over 2 l_3}e^{-2l_3t}- (2l_1+l_3)t\right\}\times\\ &&\quad {\rm Ce}_n(\om_1,q)\, {\rm ce}_n(\om_2,q)\, e^{i\om_3\ga},\quad n=0,1,2,\dots,\\ &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}+\lambda_3 {c_3^{-2}\over 2 l_3}e^{-2l_3t}- (2l_1+l_3)t\right\}\times\\ &&\quad {\rm Se}_n(\om_1,q)\, {\rm se}_n(\om_2,q)\, e^{i\om_3\ga},\quad n=1,2,3,\dots, \end{eqnarray*} where ${\rm ce}_n, {\rm se}_n$ are the even and odd Mathieu functions, ${\rm Ce}_n, {\rm Se}_n$ are the even and odd modified Mathieu functions \cite{erd,ars} and $\la_1=-4qa^2, \la_2=2q+c_n, \la_3=-\ga^2,$ and $c_n$ are eigenvalues of the Mathieu functions.} \item{Spherical coordinates \begin{eqnarray*} &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}-3l_1t\right\}\times\\ &&\quad \om_1^{1/2}J_{\pm (n+1/2)}(\al/\om_1)\, P^{\pm m}_n({\rm tanh}\, \om_2)e^{i\om_3m}, \end{eqnarray*} where $J_\nu$ is the Bessel function, $P^m_n$ is the Legendre function \cite{erd} and $\la_1=-\al^2, \la_2=-n(n+1), \la_3=-m^2.$} \item{Prolate spheroidal coordinates \begin{eqnarray*} &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}-3l_1t\right\}\times\\ &&\quad {\rm Ps}_n^{|m|}({\rm coth}\, \om_1,-a^2\la_1)\, {\rm Ps}_n^{|m|}({\rm tanh}\, \om_2,-a^2\la_1)\, e^{im\om_3}, \end{eqnarray*} where $m$ is integer, $n=0,1,2,\dots, -n\leq m\leq n, {\rm Ps}_n^m$ is the spheroidal wave function \cite{ars} and $\la_2=\la_n^{|m|}, \la_3=-m^2.$} \item{Oblate spheroidal coordinates \begin{eqnarray*} &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}-3l_1t\right\}\times\\ &&\quad {\rm Ps}_n^{|m|}(-i\tan\om_1,-a^2\la_1)\, {\rm Ps}_n^{|m|}({\rm tanh}\, \om_2,a^2\la_1)\, e^{im\om_3}, \end{eqnarray*} where $m$ is integer, $n=0,1,2,\dots, -n\leq m\leq n, {\rm Ps}_n^m$ is the spheroidal wave function and $\la_2=\la_n^{|m|}, \la_3=-m^2.$} \item{Parabolic coordinates \begin{eqnarray*} &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}-3l_1t\right\}\, e^{im\om_3}\times\\ &&\quad e^{m\om_1}\exp(\pm i\al e^{2\om_1}/2)\, _1F_1(-i\la_2/4\al+(m+1)/2,m+1,\mp i\al e^{2\om_1})\times\\ &&\quad e^{m\om_2}\exp(\pm i\al e^{2\om_2}/2)\, _1F_1(i\la_2/4\al+(m+1)/2,m+1,\mp i\al e^{2\om_2}), \end{eqnarray*} where $_1F_1$ is the confluent hypergeometric function \cite{buc,erd} and $\la_1=-\al^2, \la_3=-m^2.$} \item{Paraboloidal coordinates \begin{eqnarray*} &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}-3l_1t\right\}\, {\rm gc}_n(i\om_1;2a\al,\la_2/2\al) \times\\ &&\quad {\rm gc}_n(\om_2;2a\al,\la_2/2\al)\, {\rm gc}_n(i\om_3+\pi/2;2a\al,\la_2/2\al) \end{eqnarray*} or the same form with gc$_n$ replaced by gs$_n$. Here gc$_n$ and gs$_n$ are the even and odd nonpolynomial solutions of the Whittaker-Hill equation \cite{urw} and $n=0,1,2,\dots,$ and what is more, $\la_1=-\al^2, \la_3=\mu_n.$} \item{Ellipsoidal coordinates \[ u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}-3l_1t\right\}{\rm el}_n^m(\om_1)\, {\rm el}_n^m(\om_2)\, {\rm el}_n^m(\om_3), \] where $m$ is integer, $n=0,1,2,\dots, -n\leq m\leq n,$ el$_n^m$ is the ellipsoidal wave function \cite{ars} and $\la_1=\nu_{nm}, \la_2=\la_{nm}, \la_3=\mu_{nm}.$} \item{For the case of conical coordinates we have two types of solutions \begin{eqnarray*} &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}-3l_1t\right\}\, \om_1^{1\over 2}J_{\pm (n+{1\over 2})}(\al/\om_1)\times\\ &&\quad {\rm Ec}_n^m(\om_2)\, {\rm Ec}_n^m(\om_3),\quad n=0,1,2,\dots,\quad m=0,1,\dots,n,\\ &&u(t,\vec \om)=\exp\left\{\lambda_1 {c_1^{-2} \over 2 l_1}e^{-2l_1t}-3l_1t\right\}\, \om_1^{1\over 2}J_{\pm (n+{1\over 2})}(\al/\om_1)\times\\ &&\quad {\rm Es}_n^m(\om_2)\, {\rm Es}_n^m(\om_3),\quad n=1,2,3,\dots,\quad m=1,2,\dots,n, \end{eqnarray*} where $J_\nu$ is the Bessel function, Ec$_n^m$ and Es$_n^m$ are the even and odd Lam\'e functions \cite{erd,ars} and $\la_1=-\al^2, \la_2=-n(n+1), \la_3=-c_n^m, $ where $c_n^m$ are eigenvalues of the Lam\'e functions.} \end{enumerate} In these equations we suppose that $l_i\ne 0,\ (i=1,2,3)$. Given the condition $l_i=0$, the expressions $\exp(-2l_it)/{2l_i}$ should be replaced by $-t$. Finally, we give a list of the drift velocity vectors $\vec B(\vec x)$ providing separability of the corresponding FPEs. They have the following form: \[ \vec B (\vec x)= {\cal M}\vec x + \vec v, \] where $\vec v$ is arbitrary constant vector and ${\cal M}$ is constant matrix given by one of the following formulae: \begin{enumerate} \item ${\cal M}={\cal T} L\, {\cal T}^{-1}$, where \[ L=\left(\begin{array}{ccc} l_1 & 0 & 0\\ 0 & l_2 & 0 \\ 0 & 0 & l_3 \end{array}\right), \] $l_1,l_2,l_3$ are constants and ${\cal T}$ is an arbitrary constant $3\times 3$ orthogonal matrix, i.e. ${\cal M}$ is a real symmetric matrix with eigenvalues $l_1,l_2,l_3$. \begin{enumerate} \item $l_1,l_2,l_3$ are all distinct. The corresponding FPE has solution 1 only from the above list. The new coordinates $\om_1,\om_2,\om_3$ are given implicitly by formula \begin{equation} \label{10.1} \vec x ={\cal T} H(t)\,\vec z(\vec \om) + \vec w(t), \end{equation} where $\vec z(\vec\om)$ is given by formula 1 from (\ref{1.2}), $\vec w(t)$ is solution of system of ordinary differential equations (\ref{4.3}) and \begin{equation} \label{10.2} H(t)=\left(\begin{array}{ccc} c_1e^{l_1t} & 0 & 0\\ 0 & c_2e^{l_2t} & 0 \\ 0 & 0 & c_3e^{l_3t} \end{array}\right) \end{equation} with arbitrary constants $c_1,c_2,c_3$. \item $l_1=l_2\ne l_3$. The corresponding FPE has solutions 1--4 only from the above list. The new coordinates $\om_1,\om_2,\om_3$ are given implicitly by (\ref{10.1}), where $\vec z(\vec\om)$ is given by one of the formulae 1--4 from (\ref{1.2}) and $H(t)$ is given by (\ref{10.2}) with arbitrary constant $c_1,c_2,c_3$ satisfying the condition $c_1=c_2$ for the partially split coordinates 2--4 from (\ref{1.2}). \item $l_1=l_2=l_3$, i.e. $M=l_1 I$, where $I$ is unit matrix. The corresponding FPE has all 11 solutions, listed above. The new coordinates $\om_1,\om_2,\om_3$ are given implicitly by formula (\ref{10.1}), where $\vec z(\vec\om)$ is given by one of the eleven formulae (\ref{1.2}) and $H(t)$ is given by (\ref{10.2}) with arbitrary constants $c_1,c_2,c_3$ satisfying the condition $c_1=c_2$ for the partially split coordinates 2--4 from (\ref{1.2}) and the condition $c_1=c_2=c_3$ for the non-split coordinates 5--11 from (\ref{1.2}). \end{enumerate} \item \[ M=b\ {\cal C}_1\left(\begin{array}{ccc} 0 & \cos s & 0\\ -\cos s & 0 & \sin s \\ 0 & -\sin s & 0 \end{array}\right){\cal C}_1^{-1}+l_1I, \] where $I$ is the unit matrix and ${\cal C}_1$ is an arbitrary constant $3\times 3$ orthogonal matrix, $b,s,l_1$ are arbitrary constants and $b\ne 0$. The corresponding FPE has all 11 solutions, listed above with $l_1=l_2=l_3$. The new coordinates $\om_1,\om_2,\om_3$ are given implicitly by formula (\ref{2.7}), where $\vec z(\vec\om)$ is given by one of the eleven formulae (\ref{1.2}), ${\cal T}(t)$ is given by (\ref{3.6})--(\ref{3.7}), $\vec w(t)$ is solution of system of ordinary differential equations (\ref{4.3}) and \[ H(t)=\exp(l_1t)\left(\begin{array}{ccc} c_1 & 0 & 0\\ 0 & c_2 & 0 \\ 0 & 0 & c_3 \end{array}\right) \] with arbitrary constants $c_1,c_2,c_3$ satisfying the condition $c_1=c_2$ for the partially split coordinates 2--4 from (\ref{1.2}) and the condition $c_1=c_2=c_3$ for non-split coordinates 5--11 from (\ref{1.2}). \item \begin{eqnarray*} M&=&{\cal C}_1\left(\begin{array}{c} {1\over 2}(l_1+l_3+(l_1-l_3)\cos 2s) \\ -b\cos s \\ {1\over 2}(l_3-l_1)\sin 2s \end{array}\right.\qquad\to\\ &&\to\qquad\left.\begin{array}{cc} b\cos s & {1\over 2}(l_3-l_1)\sin 2s\\ l_1 & b\sin s \\ -b\sin s & {1\over 2}(l_1+l_3-(l_1-l_3)\cos 2s) \end{array}\right) {\cal C}_1^{-1}, \end{eqnarray*} where ${\cal C}_1$ is an arbitrary constant $3\times 3$ orthogonal matrix, $b,s,l_1,l_2$ are arbitrary constants, $l_1\ne l_3$ and $b\ne 0$. The corresponding FPE has solutions 1--4 only from the above list with $l_1=l_2\ne l_3$. The new coordinates $\om_1,\om_2,\om_3$ are given implicitly by formula (\ref{2.7}), where $\vec z(\vec\om)$ is given by one of the formulae 1--4 from (\ref{1.2}), ${\cal T}(t)$ is given by (\ref{3.6}), (\ref{3.7}) and $(iii)$ from (\ref{4a}), $\vec w(t)$ is solution of system of ordinary differential equations (\ref{4.3}) and \[ H(t)=\left(\begin{array}{ccc} c_1e^{l_1t} & 0 & 0\\ 0 & c_2e^{l_1t} & 0 \\ 0 & 0 & c_3e^{l_3t} \end{array}\right) \] with arbitrary constants $c_1,c_2,c_3$ satisfying the condition $c_1=c_2$ for the partially split coordinates 2--4 from (\ref{1.2}). \end{enumerate} Note that the above obtained solutions can be used as the basis functions to expand an arbitrary smooth solution of the equation under study in a properly chosen Hilbert space (for more details, see \cite{mil77}). The physical analysis of the obtained results seems to be very interesting, but the detailed study of this problem goes beyond the scope of the present paper. \section* {IV. R--separation of variables in the Fokker--Planck equation} In this paper, we restrict ourselves to the choice of separation Ansatz in the form (\ref{2.1}). Generally speaking, the problem of separation of variables includes the search of {\em R--separable} solutions of the more general form \cite{mil77} \begin{equation} \label{2.1a} u(t,{\vec x})=e^{R(t,{\vec x})}\vp_0(t)\prod\limits_{a=1}^3\, \vp_a\left(\omega_a(t, {\vec x}), \vec \lambda\right). \end{equation} In this case we have an analog of the system of equations (\ref{2.4c})--(\ref{2.4d}) \begin{eqnarray} && \left(2{\p R \over \p x_j}+ B_j\right){\p \omega_a\over \p x_j} + {\p\omega_a\over \p t} + \Delta \omega_a = 0,\quad a=1,2,3;\label{2.4c1}\\ && \sum\limits_{i=1}^{3}\, F_{i0}(\omega_i) {\p \omega_i\over \p x_j}{\p \omega_i\over \p x_j} + {\p R\over \p t} + \Delta R + B_a {\p R \over \p x_a}+\non\\ &&\quad + {\p R \over \p x_a}{\p R \over \p x_a} + T_0(t) + {\p B_a\over \p x_a} =0.\label{2.4d1} \end{eqnarray} Equations (\ref{2.4a})--(\ref{2.4b}) are not changed. In a way analogous to that used above we get from (\ref{2.4c1}) the form of the drift coefficients $\vec B(\vec x)$ \begin{equation} \label{3.3a} \vec B (\vec x)= {\cal M}(t)(\vec x - \vec w)+ \dot{\vec w}- 2\vec\nabla R, \end{equation} where ${\cal M}(t)$ is given by formula (\ref{m}). The compatibility conditions of the above system of PDEs (\ref{3.3a}) yield \begin{eqnarray} &&B_{1x_2}-B_{2x_1}=-2(\dot\alpha + \dot\beta \cos\gamma),\non\\ &&B_{1x_3}-B_{3x_1}=-2(\dot\beta \cos\alpha \sin\gamma - \dot\gamma \sin\alpha),\label{3.4}\\ &&B_{2x_3}-B_{3x_2}=-2(\dot\beta \sin\alpha \sin\gamma + \dot\gamma \cos\alpha).\non \end{eqnarray} As the functions $B_1, B_2, B_3$ are independent of $t$, it follows from these conditions that rot$\vec B=\vec{\rm const}$ and the functions $\alpha(t), \beta(t), \gamma(t)$ obey the system of ODE (\ref{3.5}). Thus the matrix ${\cal T}(t)$ have the form (\ref{3.6}). Consequently the following assertion holds true. \begin{theo} For the Fokker-Planck equation (\ref{1.1}) to be R--separable it is necessary that the rotor of the drift velocity vector $\vec B(\vec x)$ is constant. \end{theo} \section* {Concluding Remarks} It follows from Theorem 1 that the choice of the drift coefficients $\vec B(\vec x)$ allowing for variable separation in the corresponding FPE is very restricted. Namely, they should be linear in the spatial variables $x_1, x_2, x_3$ in order to provide separability of FPE (\ref{1.1}) into three second-order ordinary differential equations. However, if we allow for separation equations to be of lower order, then additional possibilities for variable separation in FPE arise. As an example, we give the drift coefficients \[ B_1(\vec x)=0,\quad B_2(\vec x)=0,\quad B_3(\vec x)=B_3\left(\sqrt{x_1^2+x_2^2}\right), \] where $B_3$ is arbitrary smooth function. FPE (\ref{1.1}) with these drift coefficients separates in the cylindrical coordinate system $t, \om_1=\ln \left(\sqrt{x_1^2+x_2^2}\right), \om_2=\arctan(x_1/x_2), \om_3=x_3$ into two first-order and one second-order ordinary differential equations. For the one-dimensional FPE the choice of the drift coefficients $\vec B(\vec x)$ allowing for variable separation is essentially wider \cite{lag}. \section* {Acknowledgement} I would like to thank Renat Zhdanov for his time, suggestions, and encouragement.
\section{Introduction} In a companion paper~\cite{HMPV-spde}, we discussed classical field theories subject to stochastic noise $\eta(\vec x, t)$, described by the equation \begin{equation} D \phi(\vec x,t) = F[\phi(\vec x,t)] + \eta(\vec x,t). \end{equation} Here $D$ is any linear differential operator, involving arbitrary time and space derivatives, that does {\em not} explicitly involve the field $\phi$. The function $F[\phi]$ is any forcing term, generally nonlinear in the field $\phi$. These stochastic partial differential equations (SPDEs) can be studied using a functional integral formalism which makes manifest the deep connections with quantum field theories (QFTs). Provided the noise is translation-invariant and Gaussian, it is possible to split its two-point function into an {\em amplitude} ${\cal A}$ and a {\em shape} function $g_2(x,y)$, as follows \begin{equation} G_\eta(x,y) \mathop{\stackrel{\rm def}{=}} {\cal A} \; g_2(x-y), \end{equation} with the {\em convention} that \begin{equation} \int {\mathrm d }^d \vec x\; {\mathrm d } t \; g_2^{-1}(\vec x,t) \; = \; 1 \; = \; \tilde g_2^{-1}(\vec k=\vec 0,\omega=0). \end{equation} We showed that the one-loop effective potential for homogeneous and static fields is~\cite{HMPV-spde} \begin{eqnarray} \label{E:general} {\cal V}[\phi;\phi_0] &=& {\scriptstyle{1\over2}} F^2[\phi] + {\scriptstyle{1\over2}} {\cal A} \int {{\mathrm d }^d \vec k \; {\mathrm d } \omega\over (2\pi)^{d+1}} \ln \left[ 1 + {\tilde g_2{}(\vec k,\omega) F[\phi] {\delta^2 F\over\delta\phi\;\delta\phi} \over \left( D^\dagger(\vec k,\omega) - {\delta F\over \delta\phi}^\dagger \right) \left( D(\vec k,\omega) - {\delta F\over \delta\phi} \right)} \right] - \left( \phi \to \phi_0 \right) + O( {\cal A} ^2). \end{eqnarray} Here $\phi_0$ is any convenient background field. In the absence of symmetry breaking it is most convenient to pick $\phi_0=0$, but we reserve the right to make other choices when appropriate. Equation (\ref{E:general}) is qualitatively similar to the one-loop effective potential for a self-interacting scalar QFT~\cite{Weinberg,Zinn-Justin}: \begin{eqnarray} {\cal V}[\phi;\phi_0] &=& V(\phi) + {\scriptstyle{1\over2}} \hbar \int {{\mathrm d }^d \vec k \; {\mathrm d } \omega\over (2\pi)^{d+1}} \ln \left[ 1 + { {\delta^2 V\over\delta\phi\;\delta\phi} \over \omega^2 + \vec k^2 + m^2 } \right] - \left( \phi \to \phi_0 \right) + O(\hbar^2). \end{eqnarray} The effective potential is not only a formal mathematical tool, but it also has a deep physical meaning. In fact, in previous work~\cite{HMPV-spde} we demonstrate that the effective potential for SPDEs inherits many of the physical features and information content of the effective potential for QFTs, and that searching for minima of the SPDE effective potential provides information about ``ground states'' of such SPDEs. These ground states play an important role in the study of symmetry breaking and the onset of pattern formation and structure. In this paper we apply this formalism to the specific case of the massless KPZ equation (equivalent to the massless noisy vorticity-free Burgers equation)~\cite{Frisch,KPZ,MHKZ,Sun-Plischke,Frey-Tauber} \begin{equation} \left({\partial\over\partial t} - \nu \vec\nabla^2\right) \phi = F_0+{\lambda\over2} (\vec \nabla\phi)^2 + \eta, \end{equation} and use the formalism to investigate the ultraviolet (short-distance) properties of the system. In the fluid dynamical interpretation ({\em i.e.}, the vorticity-free Burgers equation) the fluid velocity is taken to be $\vec v = - \vec\nabla \phi$. In this representation the KPZ equation is used as a model for turbulence~\cite{Frisch}, structure development in the early universe~\cite{Structure}, driven diffusion, and flame fronts~\cite{MHKZ}. In the surface growth interpretation, $\phi(\vec x,t)$ is taken to be the height of the surface (typically defined over a two-dimensional plane)~\cite{KPZ}. In this interpretation the massless KPZ equation is a natural nonlinear extension of the Edwards--Wilkinson (EW) model~\cite{Barabasi}. An explicit tadpole term ($F_0$) is included as we will soon see that it is a necessary ingredient in completing the renormalization program. (In QFTs a dynamically generated field-independent constant term in the equations of motion is often called a tadpole, though if such a term arises from the tree-level physics it is called an external current. In this KPZ context we prefer to use the word ``tadpole'' since $F_0$ will have to be renormalized, which makes the nomenclature ``external current'' inappropriate.) After renormalization, we will fix the value of the {\em renormalized} value of $F_0$ by using the symmetries of the KPZ equation. Some additional comments regarding the {\em raison-d'etre} of the tadpole may prove helpful. It is well known that both the Edwards--Wilkinson and KPZ equations may be written with or without such a constant (the tadpole) and that a finite constant term simply represents a change in the average velocity of the surface with respect to the laboratory frame of reference, and so may be chosen at will. On the other hand, as we will see below, it is crucial to include a bare tadpole in the KPZ equation in order to be able to consistently carry out the ultraviolet renormalization (at one-loop). Nevertheless, the renormalized tadpole may be chosen at will, and can be set to any convenient value after renormalization. (The convenient value we will finally adopt will be based on symmetry arguments and will be chosen to eliminate any spurious motion of the background field.) The need for a ``bare'' tadpole is not just an artifact of the effective potential renormalization but (as we will demonstrate elsewhere) is also required to carry out the one-loop ultraviolet renormalization of the full effective action. There are {\em two} important symmetries of the KPZ equation that are relevant for our analysis. The first is the symmetry under the transformation \begin{eqnarray} \phi &\to& \phi + c(t), \\ F_0 &\to& F_0 + {d c(t)\over dt}. \end{eqnarray} In the fluid dynamics interpretation this symmetry amounts to a ``gauge transformation'' of the scalar field $\phi$ that does not change the physical velocity ($\vec v = - \vec \nabla \phi$). In the surface growth interpretation this symmetry corresponds to choosing a different coordinate system that moves vertically at a speed $dc/dt$ with respect to the initial coordinate system, and so can be thought of as a type of Galilean invariance (type I) for the KPZ equation. This transformation is a symmetry of the KPZ equation for arbitrary noise. The second symmetry we consider holds under more restrictive conditions. Consider the transformation \begin{eqnarray} \vec x &\to& \vec x' = \vec x - \lambda \; \vec \epsilon \; t, \\ t &\to& t' = t, \\ \phi(\vec x,t) &\to& \phi'(\vec x',t') = \phi(\vec x,t) - \vec \epsilon \cdot \vec x. \end{eqnarray} In the fluid dynamics interpretation this symmetry is equivalent to a Galilean transformation of the fluid velocities \begin{equation} \vec v \to \vec v' = \vec v - \vec \epsilon, \end{equation} and so can also be thought of as a type of Galilean invariance (type II). In the surface growth interpretation this symmetry amounts to choosing a different coordinate system that is tilted at an angle to the vertical, with \begin{equation} \tan(\theta) = ||\,\vec \epsilon\, ||, \end{equation} and for this reason this transformation is often referred to as tilt invariance. While this type II Galilean invariance is an exact invariance of the zero-noise KPZ equation, it is important to keep in mind that once noise is added to the system, this transformation will remain a symmetry only if the noise is translation-invariant and {\em temporally white}. This can be seen by first looking at the noise two point function \begin{equation} G_\eta(\vec x_1,t_1; \vec x_2,t_2) = {\cal A} \; g_2(\vec x_1-\vec x_2, t_1-t_2), \end{equation} then considering \begin{eqnarray} \vec x_1 - \vec x_2 &\to& \vec x_1' - \vec x_2' = \vec x_1 - \vec x_2 - \lambda \; \vec \epsilon \; (t_1 - t_2), \\ t_1 - t_2 &\to& t_1' - t_2' = t_1-t_2, \end{eqnarray} and finally noting that the noise two-point function is invariant if and only if its support is limited by the constraint $t_1=t_2$, that is, \begin{equation} G_\eta(\vec x_1,t_1; \vec x_2,t_2) = {\cal A} \; \hat g_2(\vec x_1-\vec x_2) \; \delta(t_1-t_2). \end{equation} But this is the {\em definition} of translation-invariant temporally white noise. We will use these two symmetries extensively in the body of the paper: the type I Galilean invariance is used to guarantee that the background field $\phi_0$ is kept stationary, even in the presence of interactions and nonlinearities, and the type II Galilean invariance is similarly used to guarantee that the average slope of the background field is zero. It is best (in fact essential) to do this only after the renormalization program has been completed, so for the time being we will explicitly keep track of both the tadpole term $F_0$ and the background field $\phi_0$. Applying the formalism developed in~\cite{HMPV-spde} to the massless KPZ equation, we demonstrate that the effective potential is one-loop ultraviolet renormalizable in 1, 2, and 3 space dimensions (and is not ultraviolet renormalizable in 4 or higher space dimensions). We will discover a formal relationship between the one-loop effective potential for the massless KPZ equation in $d+1$ dimensions ($d$ space and 1 time dimensions) and that of the massless $\lambda \phi^4$ QFT in $d+2$ Euclidean dimensions, and show that the massless KPZ equation exhibits many of the properties seen in massless $\lambda \phi^4$ QFT. In particular, we will see that the stochastic system undergoes dynamical symmetry breaking (DSB) in 1 and 2 space dimensions, and no symmetry breaking in 3 dimensions. This DSB is due to an analog of the Coleman--Weinberg mechanism of QFT. In 2 space dimensions the presence of a short-distance (ultraviolet) logarithmic divergence implies the running of the coupling constant $\lambda$ with the energy scale, and the existence of a non-zero beta function for this coupling. We feel that the unexpected presence of DSB in 1 and 2 space dimensions is a matter of deep importance, and is something that would be very difficult to deduce by any other means. \section{Effective Potential: massless KPZ} The massless KPZ equation (massless noisy vorticity-free Burgers equation) is~\cite{Frisch,KPZ,MHKZ,Sun-Plischke,Frey-Tauber} \begin{equation} \label{E:KPZ} \left({\partial\over\partial t} - \nu \vec\nabla^2\right) \phi = F_0 +{\lambda\over2} (\vec\nabla\phi)^2 + \eta. \end{equation} We have introduced an explicit ``bare'' tadpole term $F_0$ as remarked above. If we now restrict attention to homogeneous and static fields, $\phi(\vec x,t) \to \phi_{ homogeneous-static}$, and $\phi_0(\vec x,t) \to (\phi_0)_{ homogeneous-static}$, then from equation (\ref{E:general}) it is easy to see that ${\cal V}[\phi,\phi_0] \equiv 0$, and so the one-loop effective potential is uninteresting. In order to see this, note that for a homogeneous static field configuration \begin{equation} F[\phi] \to F_0; \qquad {\delta F\over\delta\phi(x)} \to 0; \qquad {\delta^2 F \over \delta\phi(x) \delta\phi(y)} = + \lambda \vec \nabla_x \cdot \vec \nabla_y \delta^d (\vec x, \vec y) \to - \lambda \vec \nabla_x^2 \to + \lambda \vec k^2. \end{equation} Thus the integrand appearing in (\ref{E:general}) is independent of the field $\phi$, and ${\cal V}[\phi;\phi_0]$ is zero as asserted. (This simplification does not hold for the massive noisy vorticity-free Burgers equation [massive KPZ equation] where the driving force is replaced by $F[\phi] \to F[\phi] - m^2 \phi$. We have calculated ${\cal V}[\phi;\phi_0]$ explicitly for this system and shown it is non-zero. We do not report the details here, as this is a simple exercise and the result is of limited physical interest.) What {\em is} physically interesting on the other hand, is to study the {\em massless} KPZ equation and to consider a linear and static field configuration \begin{equation} \phi = - \vec v \cdot \vec x, \end{equation} where $\vec v$ is now a constant vector. Notice that for this choice one has $D \phi = 0$. In the hydrodynamic interpretation of the massless KPZ equation this corresponds to a constant velocity flow: $\vec v = - \vec\nabla\phi$. In the surface growth interpretation, $\vert \vert \, \vec v \, \vert \vert$ corresponds to a constant slope of the surface~\cite{KPZ}. There is an instructive analogy with QED here: taking a constant vector potential in QED is relatively uninteresting, it corresponds to zero electromagnetic field strength and can be gauged away (modulo topological constraints). On the other hand, a constant electromagnetic field strength can be described by a linear vector potential, and leads to such useful quantities as the Euler--Heisenberg effective potential and the Schwinger effective Lagrangian for QED~\cite{Weinberg,Zinn-Justin}. Inspection of the derivation contained in reference~\cite{HMPV-spde} reveals that although the effective potential (\ref{E:general}) was originally defined for homogeneous and static fields, for the massless KPZ equation (\ref{E:KPZ}), it continues to make sense for linear and static fields. The fact that for the massless KPZ equation $F[\phi]$ is position independent for these linear static fields is essential to this observation. For such a field configuration, $\phi = - \vec v \cdot \vec x$, we get \begin{equation} F[\phi] \to F_0 + {\lambda\over2} v^2 ; \qquad {\delta F\over\delta\phi(x)} \to -\lambda \vec v \cdot \vec \nabla_x ; \qquad {\delta^2 F \over \delta\phi(x) \delta\phi(y)} = + \lambda \vec \nabla_x \cdot \vec \nabla_y \to - \lambda \vec \nabla_x^2 \to + \lambda \vec k^2. \end{equation} The zero-loop effective potential is \begin{equation} {\cal V}_{ zero-loop}[v;v_0] = {\scriptstyle{1\over2}} \left[ \left( F_0 + {\scriptstyle{1\over2}} {\lambda} v^2 \right)^2 - \left( F_0 + {\scriptstyle{1\over2}} {\lambda} v_0^2 \right)^2 \right], \end{equation} where the background field is $\phi_0=-\vec v_0 \cdot \vec x$. Note that this zero-loop effective potential is formally equivalent to that of $\lambda \phi^4$ QFT---with the velocity $v$ playing the role of the quantum field $\phi_{ QFT}$. Even at zero loops (tree level) we see that if we were to have $F_0<0$ the effective potential would take on the ``Mexican hat'' form, so that the onset of spontaneous symmetry breaking (SSB) would not be at all unexpected~\cite{Weinberg,Zinn-Justin}. However, as previously mentioned, the renormalized value of $F_0$ is not physically relevant and can always be changed by a type I Galilean transformation. We will soon see that SSB is not a feature of the KPZ equation. Instead we encounter a much more subtle effect: the onset of dynamical symmetry breaking (DSB) which we have detected via a one-loop computation. We start the one-loop calculation by noting that \begin{equation} D - {\delta F\over\delta \phi} = \partial_t + \lambda \; \vec v \cdot \vec \nabla - \nu \vec \nabla^2 \qquad \to \qquad -i\omega +i\lambda \vec v \cdot \vec k + \nu \vec k^2, \end{equation} while for the adjoint quantity \begin{equation} D^\dagger - {\delta F\over\delta \phi}^\dagger = -\partial_t - \lambda \; \vec v \cdot \vec \nabla - \nu \vec \nabla^2 \qquad \to \qquad +i\omega -i\lambda \vec v \cdot \vec k + \nu \vec k^2, \end{equation} so that \begin{equation} \left(D^\dagger -{\delta F\over \delta\phi}^\dagger\right) \left(D - {\delta F \over\delta\phi}\right) = -(\partial_t + \lambda \vec v \cdot \vec \nabla)^2 + \nu^2 (\vec \nabla^2)^2 \qquad \to \qquad (\omega - \lambda \; \vec v \cdot \vec k)^2 + \nu^2 (\vec k^2)^2. \end{equation} Using this we specialize equation (\ref{E:general}) to \begin{eqnarray} {\cal V}[v;v_0] &=& {\scriptstyle{1\over2}} (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2)^2 + {\scriptstyle{1\over2}} {\cal A} \int {{\mathrm d }^d \vec k \; {\mathrm d } \omega\over (2\pi)^{d+1}} \ln \left[ 1 + {\tilde g_2{}(\vec k,\omega) \lambda (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2) \vec k^2 \over (\omega - \lambda \vec v \cdot \vec k)^2 + \nu^2 (\vec k^2)^2} \right] - \left( \vec v \to \vec v_0 \right) + O( {\cal A} ^2). \end{eqnarray} Equivalently \begin{eqnarray} {\cal V}[v;v_0] &=& {\scriptstyle{1\over2}} (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2)^2 + {\scriptstyle{1\over2}} {\cal A} \int {{\mathrm d }^d \vec k \; {\mathrm d } \omega\over (2\pi)^{d+1}} \ln \left[ { (\omega - \lambda \vec v \cdot \vec k)^2 + \nu^2 (\vec k^2)^2 + \tilde g_2{}(\vec k,\omega) \lambda (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2) \vec k^2 \over (\omega- \lambda \vec v \cdot \vec k)^2 + \nu^2 (\vec k^2)^2} \right] \nonumber\\ && - \left( \vec v \to \vec v_0 \right) + O( {\cal A} ^2). \end{eqnarray} This is as far as we can go {\em without making further assumptions about the noise}. For instance, one very popular choice is {\em temporally white}, which means delta function correlated in time so that $\tilde g_2(\vec k,\omega)\to \tilde g_2(\vec k)$ is a function of $\vec k$ only. We can shift the integration variable $\omega$ to $\omega - \lambda \vec v \cdot \vec k$. (The frequency integral is convergent.) The resulting integral becomes \begin{eqnarray} {\cal V}[v;v_0] &=& {\scriptstyle{1\over2}} (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2)^2 + {\scriptstyle{1\over2}} {\cal A} \int {{\mathrm d }^d \vec k \; {\mathrm d } \omega\over (2\pi)^{d+1}} \ln \left[ {\omega^2 + \nu^2 (\vec k^2)^2 +\tilde g_2{}(\vec k) \lambda (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2) \vec k^2 \over \omega^2 + \nu^2 (\vec k^2)^2 } \right] \nonumber\\ && - \left( \vec v \to \vec v_0 \right) + O( {\cal A} ^2). \end{eqnarray} The fact that after this change of variables the denominator of the logarithm in the two integrands is independent of the fields $\vec v$ and $\vec v_0$ is actually a surprisingly deep result, related to the fact that the Jacobian functional determinant encountered in~\cite{HMPV-spde} is a field-independent constant for the KPZ equation. We now make use of the integral \begin{equation} \int_{-\infty}^{+\infty} {\mathrm d } \omega \ln \left( {\omega^2 + X^2\over \omega^2 + Y^2} \right) = 2\pi \left( X - Y \right), \end{equation} to deduce \begin{eqnarray} {\cal V}[v;v_0] &=& {\scriptstyle{1\over2}} \left[ (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2)^2 - (F_0+ {\scriptstyle{1\over2}} {\lambda} v_0^2)^2 \right] \nonumber\\ && + {\scriptstyle{1\over2}} {\cal A} \int {{\mathrm d }^d \vec k \over (2\pi)^{d}} \left\{ \sqrt{ \nu^2 (\vec k^2)^2 + \tilde g_2{}(\vec k) \lambda (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2) \vec k^2 } - \sqrt{ \nu^2 (\vec k^2)^2 + \tilde g_2{}(\vec k) \lambda (F_0+ {\scriptstyle{1\over2}} {\lambda} v_0^2) \vec k^2} \right\} \nonumber\\ && + O( {\cal A} ^2). \end{eqnarray} This may be simplified by extracting an explicit factor of $\nu^2 \vec k^2$. If we do so, the one-loop effective potential becomes \begin{eqnarray} \label{E:eff-potl} {\cal V}[v;v_0] &=& {\scriptstyle{1\over2}} \left[ (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2)^2 - (F_0+ {\scriptstyle{1\over2}} {\lambda} v_0^2)^2 \right] \nonumber\\ && + {\scriptstyle{1\over2}} {\cal A} \nu \int {{\mathrm d }^d \vec k |k| \over (2\pi)^{d}} \left\{ \sqrt{ \vec k^2 + \tilde g_2{}(\vec k) {\lambda\over\nu^2} (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2) } - \sqrt{ \vec k^2 + \tilde g_2{}(\vec k) {\lambda\over\nu^2} (F_0+ {\scriptstyle{1\over2}} {\lambda} v_0^2) } \right\} \nonumber\\ && + O( {\cal A} ^2). \end{eqnarray} Remembering that $\lim_{\vec k \rightarrow \vec 0} {\tilde g_2{}}(\vec k) = 1$, it is clear that there are no infrared divergences $(\vec k \rightarrow \vec 0)$ to worry about, at least for this effective potential at one loop order. Furthermore, modulo possibly perverse choices for the spatial noise spectrum, the tadpole term $F_0$ is essential in renormalizing the theory. Note that the bare potential contains terms proportional to $v^0$, $v^2$, and $v^4$, whereas the one-loop contribution, when one expands it in powers of $v^2$ has terms proportional to $v^{2n}$ for $n=1,2,3...$. For definiteness, let us now take the spatial noise spectrum to be cutoff white (the temporal spectrum has already been chosen to be exactly white), {\em i.e.}, \begin{equation} \tilde g_2(\vec k) =\tilde g_2(\vert \vec k \vert) = \Theta(\Lambda - k). \end{equation} The effective potential is then \begin{eqnarray} \label{E:effpotential} {\cal V}[v;v_0] &=& {\scriptstyle{1\over2}} \left[ (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2)^2 - (F_0+ {\scriptstyle{1\over2}} {\lambda} v_0^2)^2 \right] \nonumber\\ && + {\scriptstyle{1\over2}} {\cal A} \nu \int_{k < \Lambda} {{\mathrm d }^d \vec k |k| \over (2\pi)^{d}} \left\{ \sqrt{ \vec k^2 + {\lambda\over\nu^2} (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2) } - \sqrt{ \vec k^2 + {\lambda\over\nu^2} (F_0+ {\scriptstyle{1\over2}} {\lambda} v_0^2) } \right\} \nonumber\\ && + O( {\cal A} ^2). \end{eqnarray} With this choice of noise, the $v^2$ term is ultraviolet divergent and proportional to $\Lambda^d$, the $v^4$ term is proportional to $\Lambda^{d-2}$, and the $v^6$ term to $\Lambda^{d-4}$. To have any hope of absorbing the infinities into the bare action, thereby permitting us to take the $\Lambda\to\infty$ limit, we must have $d<4$ (because there is no $v^6$ term in the bare potential). That is: the massless KPZ equation (subject to white noise) is one-loop ultraviolet renormalizable {\em only} in 1, 2, and 3 space dimensions. (In $0$ space dimensions the KPZ equation is trivial.) And even so, one-loop renormalizability requires an explicit tadpole term. (Without a tadpole term there is no term proportional to $v^2$ in the zero-loop potential, and therefore there is no possibility of renormalizing the leading divergence.) Strictly speaking the claim of one-loop renormalizability also requires investigation of the wave-function renormalization. This appears in the one-loop effective action, which is beyond the scope of the present paper (here we confine attention to the one-loop effective potential), and will be discussed in future work. It is also clear from the above that the ultraviolet renormalizability of the KPZ equation depends critically on the high momentum behaviour of the noise. Let us temporarily return to equation (\ref{E:eff-potl}), and suppose that the noise is power-law distributed in the ultraviolet with $\tilde g_2(\vec k) = \tilde g_2(\vert \vec k \vert) \approx (k_0/k)^{\theta} \; \Theta(\Lambda-k)$. Then the $n$'th term in the expansion has ultraviolet behaviour proportional to $(F_0+ {\scriptstyle{1\over2}} \lambda v^2)^{n} \Lambda^{d+2-2n-n \theta}$. The massless KPZ equation is then one-loop ultraviolet renormalizable provided the $n=3$ term (and higher terms) are ultraviolet finite, that is, for $d<4+3\theta$. We will not deal with these issues any further in this paper but will instead confine ourselves to white noise. (Recall that it is only at intermediate stages that the white noise is regulated by an ultraviolet spatial cutoff. After renormalization, we will be considering noise that is exactly white.) We can also point out an unexpected result of this calculation: There is a formal connection between the massless KPZ equation and massless $\lambda (\phi^4)_{[d+1]+1}$ QFT. Consider equation (\ref{E:effpotential}). This is recognizable (either from QFT or equilibrium statistical field theory) as the effective potential for $\lambda \phi^4$ Lorentzian QFT in $d+1$ space dimensions, or equivalently $\lambda\phi^4$ statistical field theory in $(d+1)+1$ Euclidean spacetime dimensions~\cite{Weinberg,Zinn-Justin}. To make the connection, interpret $\lambda F_0/\nu^2$ as the mass term $m^2$ of the QFT, $\lambda^2/\nu^2$ as $\lambda_{ QFT}$ (the coupling constant of the QFT), and $v$ as the mean field $\phi_{ QFT}$. Note that we have at this stage (temporarily) kept the parameters $(F_0)_{ renormalized}$ and $v_0$ non-zero for clarity. We now invoke the type I and type II Galilean symmetries of the KPZ equation to fix these parameters: \noindent---(1) In the fluid dynamical interpretation, the tadpole term does not have any physical significance. (Because of the symmetry $F_0\to F_0-\kappa$, $\phi\to\phi-\kappa t$, the tadpole does not affect any of the physical fluid velocities.) We are interested in a static background field configuration $\vec v_0$, but from the equation of motion we see that for the spatially averaged field \begin{equation} {1\over \Omega} \left\langle \int {\partial \phi \over \partial t} \; {\mathrm d }^d\vec x \right\rangle = F_0 + {\scriptstyle{1\over2}} \lambda v_0^2 + O( {\cal A} ), \end{equation} so we must set the {\em renormalized} value of $F_0$ to $- {\scriptstyle{1\over2}} \lambda v_0^2$. (This works for arbitrary Gaussian noise.) Secondly, the type II Galilean invariance of the fluid dynamical system allows us to set $\vec v_0 = \vec 0$ without loss of generality: pick a coordinate system moving with the bulk fluid velocity of the background field. (This requires translation-invariant and temporally white noise.) Combining the two symmetries yields $(F_0)_{ renormalized}=0$ and $v_0=0$. \noindent---(2) In the surface growth interpretation $(F_0)_{ renormalized}$ is a contribution to the spatially-averaged velocity of the interface. This average velocity term can always be scaled away by the shift $\phi \rightarrow \phi - (F_0)_{ renormalized} \; t$, which in the surface growth interpretation results from a type I Galilean transformation that places one in an inertial frame moving with the average surface profile. Doing so again sets the {\em renormalized} value of $(F_0)_{ renormalized}$ to $- {\scriptstyle{1\over2}} \lambda v_0^2$. Secondly, the type II Galilean invariance now corresponds to tilting the coordinate system away from the vertical. For any constant slope background field a simple tilt can then be used to set the slope to zero. Combining the two symmetries yields $(F_0)_{ renormalized}=0$ and $v_0=0$, so the surface growth interpretation and the fluid dynamics interpretation are in agreement as to what the physically relevant choice of parameters is. The perhaps somewhat unusual way in which the tadpole $F_0$ is first introduced and then renormalized to zero has a direct analog in ordinary $\lambda \phi^4$ QFT: when we consider {\em massless} $\lambda \phi^4$ QFT it is well known that a bare mass must be introduced to successfully complete the renormalization program. It is only {\em after} renormalization that the renormalized mass may be set to zero, and massless $\lambda \phi^4$ QFT makes sense only in this post-renormalization fashion~\cite{Weinberg,Zinn-Justin}. \subsection{Massless KPZ: $d=1$} For $d=1$ we are interested in the integral \begin{equation} \int_0^{\Lambda} {\mathrm d } k^2 \left[ \sqrt{k^2+ a}- \sqrt{k^2 + b} \right]. \end{equation} This integral is the restriction to one spatial dimension of (\ref{E:effpotential}). It is divergent, and it will require renormalization to extract physical answers from this formal result. The most direct way of proceeding is via the ``differentiate and integrate'' trick where one considers~\cite{Weinberg} \begin{equation} {\cal I}(a) \mathop{\stackrel{\rm def}{=}} \int_0^{\infty} {\mathrm d } k^2 \left[ \sqrt{k^2+ a}- \sqrt{k^2} \right]. \end{equation} Now differentiate twice \begin{equation} {{\mathrm d }^2 \;{\cal I}(a) \over {\mathrm d } a^2} = -{1\over4}\int_0^{\infty} {\mathrm d } k^2 {1\over({k^2+ a })^{3/2}} = -{1\over 2 \sqrt{a}}. \end{equation} (This last integral is now well defined and finite.) If we integrate the above equation twice, we get \begin{equation} {\cal I}(a) = \kappa a - {2\over3} a^{3/2}. \end{equation} Here $\kappa$ is a constant of integration (which happens to be infinite). There would, in principle, be a second constant of integration, but that is fixed to be zero by the condition that ${\cal I}(0) =0$. We absorb $\kappa$ into the bare action, where it renormalizes $F_0$. That is, we write \begin{equation} (F_0)_{ bare} = (F_0)_{ renormalized} + {\cal A} \;[\delta F_0] + O( {\cal A} ^2), \end{equation} where the counterterm $[\delta F_0]$ is chosen so as to render the integral (when expressed in terms of renormalized quantities) finite~\cite{Weinberg,Zinn-Justin}. A brief calculation yields \begin{equation} \delta F_0 = -\frac{\kappa \lambda}{4 \pi \nu}. \end{equation} In one space dimension the only parameter that gets renormalized at order $O( {\cal A} )$ is this tadpole term. In terms of renormalized quantities the effective potential is thus \begin{eqnarray} \label{E:KPZ-1} {\cal V}[v;v_0;d=1] &=& {\scriptstyle{1\over2}} \left[ (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2)^2 - (F_0+ {\scriptstyle{1\over2}} {\lambda} v_0^2)^2 \right] - {1\over6\pi} {\cal A} {\lambda^{3/2} \over \nu^2} \left[ \left(F_0+ {\scriptstyle{1\over2}} {\lambda} v^2\right)^{3/2} - \left(F_0+ {\scriptstyle{1\over2}} {\lambda} v_0^2\right)^{3/2} \right] + O( {\cal A} ^2). \end{eqnarray} These are now renormalized parameters at order $O( {\cal A} )$. Setting the slope $v_0$ and the renormalized tadpole $F_0$ to their physical values of zero yields \begin{eqnarray} {\cal V}[v;v_0=0;d=1] &=& {\lambda^2\over8}v^4 - {1\over6\pi} {\cal A} {\lambda^3 \over {2^{{3}/{2}} \nu^2}} |v|^3 + O( {\cal A} ^2). \end{eqnarray} Note that the potential is not analytic at zero field (a phenomenon well-known from massless QFTs) \cite{Weinberg,Zinn-Justin}. For large $v$ the classical potential dominates, while for small $v$ the one-loop effects dominate. The minimum of the potential is not at $v=0$, since the $-|v|^3$ term is negative and dominant near $v=0$. Thus we encounter something very interesting ---the system undergoes {\em dynamical symmetry breaking} (DSB) in a manner qualitatively similar to the Coleman--Weinberg mechanism of particle physics. The qualitative form of the effective potential is sketched in figure~\ref{F:d1}. \begin{figure}[htb] \vbox{\hfil\epsfbox{kpz-1d.eps}\hfil} \caption{% The one-loop effective potential for the KPZ equation in $d=1$ space dimensions. The behaviour at the origin is non-analytic in that the third derivative is discontinuous. This distorted Mexican hat potential indicates the onset of dynamical symmetry breaking. For comparison we also plot the zero-loop (``classical'') effective potential. Note that there is no symmetry breaking at tree level. }\label{F:d1} \end{figure} Symmetry breaking is said to be spontaneous if there is a symmetry in the potential that is not shared by the zero-loop ground states ({\em e.g.}, the Higgs mechanism). If the symmetry is preserved at the classical level (zero-noise), but is broken once fluctuations are taken into account (broken by loop effects) then the symmetry breaking is said to be dynamical ({\em e.g.}, massless $\lambda \phi^4$ theory, Coleman--Weinberg mechanism). {For} small $\vert \vert \, \vec v \, \vert \vert$ the one-loop contributions drive the minimum of the potential away from zero field. To find the location of the minimum we calculate \begin{equation} {{\mathrm d }{\cal V}[v;v_0=0;d=1]\over {\mathrm d } v} = {\scriptstyle{1\over2}} \lambda^2 v^3 -{1\over2\pi}\; {\cal A} \;{\lambda^3 \over {2^{{3}/{2}} \nu^2}} \hbox{sign} (v)\, v^2 + O( {\cal A} ^2) = 0. \end{equation} This permits us to {\em estimate} the shift in the expectation value of the velocity field \begin{equation} v_{ min} = \pm {\cal A} \; {\lambda\over2\pi \; 2^{1/2} \; \nu^2} + O( {\cal A} ^2). \end{equation} Unfortunately, the presence of the unknown $O( {\cal A} ^2)$ terms renders it impossible to make any definitive statement about the precise value of $v_{ min}$, apart from the fact that it is non-zero~\cite{Rivers,Pokorsky}. (This is a common feature in DSB, as perturbatively detecting the occurrence of DSB is easier than finding the precise location of the minimum.) To complete the specification of $v_{ min}$ one would have to calculate the $O( {\cal A} ^2)$ terms in the effective potential and verify that the one-loop estimate of $v_{ min}$ occurs at values of the velocity where the $O( {\cal A} ^2)$ term is negligible. This is not a trivial task, and we refer the reader to several texts where this is more fully addressed~\cite{Rivers,Pokorsky}. This DSB is particularly intriguing in that it suggests the possibility of a noise driven pump. For example, in thin pipes where the flow is essentially one-dimensional, and provided the physical situation justifies the use of the vorticity-free Burgers equation for the fluid, this result indicates the presence of a bimodal instability leading to the onset of a bulk fluid flow with velocity dependent on the noise amplitude. In the surface growth (line growth) interpretation the onset of DSB corresponds to an initially flat line breaking up into a sawtooth pattern of domains in which the slope takes on the values $\pm v_{ min}$. \subsection{Massless KPZ: $d=2$} The present discussion is relevant to either (1) surface evolution on a two dimensional substrate, or (2) thin superfluid films (since superfluids are automatically vorticity-free, justifying the application of the zero-vorticity Burgers equation). An immediate consequence of the analogy between the one-loop effective potential for the massless KPZ equation (for white noise) and that for the massless $\lambda \phi^4$ QFT is that we can write down the renormalized one-loop effective potential for $d=2$ (space dimensions) by inspection, merely by recalling that for the $d=4$ (spacetime dimensions) scalar field theory we have (see, {\em e.g.,} \cite{Weinberg,Zinn-Justin}) \begin{eqnarray} {\cal V}[v;v_0;d=2] &=& {\scriptstyle{1\over2}} \left\{ [F_0(\mu)+ {\scriptstyle{1\over2}} {\lambda(\mu)} v^2]^2 - [F_0(\mu)+ {\scriptstyle{1\over2}} {\lambda(\mu)} v_0^2]^2 \right\} \nonumber\\ && + {\scriptstyle{1\over2}} {\cal A} {1 \over (2\pi)^{2}} {\lambda^2\over\nu^3} \Bigg\{ [F_0(\mu)+ {\scriptstyle{1\over2}} {\lambda(\mu)} v^2]^2 \ln\left[{F_0(\mu)+ {\scriptstyle{1\over2}} {\lambda(\mu)} v^2\over \mu^2 }\right] \nonumber\\ && \qquad\qquad\qquad -[F_0(\mu)+ {\scriptstyle{1\over2}} {\lambda(\mu)} v_0^2]^2 \ln\left[{F_0(\mu)+ {\scriptstyle{1\over2}} {\lambda(\mu)} v_0^2\over \mu^2 }\right] \Bigg\} + O( {\cal A} ^2). \end{eqnarray} Here $\mu$ is the renormalization scale, as normally used in QFT \cite{Weinberg,Zinn-Justin}. (In a condensed matter setting this might be thought of as a measure of the coarse-graining scale.) Its presence is a side effect of the logarithmic divergences, which happen to occur in $d=2$ space dimensions for the KPZ equation. If we now tune $v_0$ and the renormalized value of the tadpole $F_0$ to their physical values of zero we obtain \begin{eqnarray} {\cal V}[v;v_0=0;d=2] &=& {\lambda^2\over 8} v^4 + {\scriptstyle{1\over2}} {\cal A} {1 \over (2\pi)^{2}} {\lambda^4\over4\nu^3} v^4 \ln\left({v^2\over \mu^2 }\right) + O( {\cal A} ^2). \end{eqnarray} This potential is zero at $v=0$, then becomes negative, though for large enough fields ($v>\mu$) the potential again becomes positive. That this potential has a nontrivial minimum exhibiting DSB is exactly the analog in this massless KPZ problem of the well-known Coleman--Weinberg mechanism encountered in the extensions of the standard model of particle physics~\cite{Weinberg,Zinn-Justin}. As in $d=1$, it is easy to see that the minimum effective potential occurs for $v_{ min} \neq 0$, but because of the presence of unknown $O( {\cal A} ^2)$ terms it is difficult to give a good estimate for the value of $v_{ min}$~\cite{Rivers,Pokorsky}. The qualitative form of the effective potential is sketched in figure~\ref{F:d2}. If we estimate the location of the minimum by differentiating the effective potential we obtain \begin{equation} v_{ min} = \pm \mu \; \exp\left( -{(2\pi)^2\;\nu^3\over2\;\lambda^2 \; {\cal A} } - {1\over4} +O( {\cal A} ) \right). \end{equation} Note that $v_{min} \to 0$ as $ {\cal A} \to 0$, as it should, to recover the tree level minimum $v_{min}=0$. \begin{figure}[htb] \vbox{\hfil\epsfbox{kpz-2d.eps}\hfil} \caption{% The one-loop effective potential for the KPZ equation in $d=2$ space dimensions. The behaviour at the origin is non-analytic in that the fourth derivative exhibits a logarithmic singularity. This distorted Mexican hat potential indicates the onset of dynamical symmetry breaking. For comparison we also plot the zero-loop (``classical'') effective potential, note that there is no symmetry breaking at tree level. }\label{F:d2} \end{figure} Following the analysis of~\cite{Gato} we can immediately extract the one-loop beta function. In order to do so, we make use of the fact that the bare effective potential does not depend on the renormalization scale: \begin{equation} {\mu \; {\mathrm d } \over {\mathrm d }\mu} \; {\cal V}[v;v_0;d] =0. \end{equation} We get \begin{equation} \label{E:beta1} {\mu \; {\mathrm d } \over {\mathrm d }\mu} \; \left( {\lambda^2(\mu)\over 8} v^4 \right) = { {\cal A} \over16\pi^2} {\lambda^4 v^4\over\nu^3} + O( {\cal A} )^2. \end{equation} Comparing the coefficients of the $v^4$ terms we obtain \begin{equation} \label{E:beta2} \beta_\lambda \mathop{\stackrel{\rm def}{=}} {\mu \; {\mathrm d } \over {\mathrm d }\mu} \; \lambda = { {\cal A} \over4\pi^2} {\lambda^3\over\nu^3} + O( {\cal A} )^2. \end{equation} We cannot extract the beta function for the wavefunction renormalization of $\phi$ (or $v$) from the present analysis. This would require a calculation of the effective action for an inhomogeneous field, an issue which we postpone for the future. Equations (\ref{E:beta1}) and (\ref{E:beta2}) are correct because one can show~\cite{HMPV-kernel} that to one loop there is no wavefunction renormalization for the KPZ field in this background. \subsection{Massless KPZ: $d=3$} This case is of physical interest for three-dimensional fluids. For $d=3$ we are interested in the integral \begin{equation} \int_0^\infty {\mathrm d } k^2 k^2 \left[ \sqrt{k^2+ a }- \sqrt{k^2 + b} \right] . \end{equation} Define the quantity \begin{equation} {\cal I}(a) \mathop{\stackrel{\rm def}{=}} \int_0^{\infty} {\mathrm d } k^2 k^2 \left[ \sqrt{k^2+ a}- \sqrt{k^2} \right]. \end{equation} The ``differentiate and integrate'' trick~\cite{Weinberg} leads to \begin{equation} {{\mathrm d }^3 \;{\cal I}(a) \over {\mathrm d } a^3} = {3\over8}\int_0^{\infty} {\mathrm d } k^2 {k^2\over({k^2+ a})^{5/2}} = {1\over 2 \sqrt{a}}. \end{equation} (The integral is now well behaved and finite). We now integrate this thrice to obtain \begin{equation} {\cal I}(a) = \kappa_1 a + \kappa_2 a^2 + {4\over15} a^{5/2}. \end{equation} Here $\kappa_1$ and $\kappa_2$ are now two constants of integration (which happen to be infinite). There would in principle be a third constant of integration, but that is fixed to be zero by the condition that ${\cal I}(0) =0$. (The sign in front of the $a^{5/2}$ term is important, since this sign is positive we will see that there is no possibility of DSB in three space dimensions.) We absorb $\kappa_1$ and $\kappa_2$ into the bare potential, where they renormalize both $F_0$ and $\lambda$. The relevant counterterms are \begin{equation} \delta F_0 = -{\kappa_1\over8\pi^2} {\lambda\over\nu}\; , \; \; \; \; \; {\rm and} \; \; \; \; \; \delta \lambda = - {\kappa_2\over4\pi^2} \left({\lambda\over\nu}\right)^3 \; , \end{equation} so as to yield \begin{eqnarray} {\cal V}[v;v_0;d=3] &=& {\scriptstyle{1\over2}} \left[ (F_0+ {\scriptstyle{1\over2}} {\lambda} v^2)^2 - (F_0+ {\scriptstyle{1\over2}} {\lambda} v_0^2)^2 \right] + {1\over 30\pi^2} {\cal A} {\lambda^{5/2} \over \nu^4} \left[ \left( F_0+ {\scriptstyle{1\over2}} {\lambda} v^2\right)^{5/2} - \left( F_0+ {\scriptstyle{1\over2}} {\lambda} v_0^2 \right)^{5/2} \right] \nonumber\\ && + O( {\cal A} ^2). \end{eqnarray} These are now all renormalized parameters at order $O( {\cal A} )$. Setting to zero both $v_0$ and the renormalized value of $F_0$ we have \begin{eqnarray} {\cal V}[v;v_0=0;d=3] &=& {\lambda^2\over8} v^4 + {1\over 30\pi^2} {\cal A} {\lambda^{5} \over 2^{{5}/{2}} \nu^4} |v|^5 + O( {\cal A} ^2). \end{eqnarray} At zero-loops the vacuum is symmetric at $v=0$. Adding one-loop physics does not change this. Note that there is an all-important sign change in the one-loop contribution to the effective potential in comparing $d=3$ with $d=1$. The DSB that is so interesting in $d=1$, and via the Coleman--Weinberg mechanism in $d=2$, is now completely absent in $d=3$. Observe also that the effective potential is non-analytic at $v=0$. The qualitative form of the effective potential is sketched in figure~\ref{F:d3}. \begin{figure}[htb] \vbox{\hfil\epsfbox{kpz-3d.eps}\hfil} \caption{% The one-loop effective potential for the KPZ equation in $d=3$ space dimensions. The behaviour at the origin is non-analytic in that the fifth derivative is discontinuous. The fact that the potential has only one minimum indicates that there is no symmetry breaking. The zero-loop (``classical'') potential is also plotted for comparison purposes. In $d=3$ space dimensions one-loop effects are not dramatic, and the qualitative form of the effective potential is not significantly altered by one loop effects. }\label{F:d3} \end{figure} \section{Discussion} In this paper we have explicitly calculated the renormalized one-loop effective potential for the massless KPZ equation in 1, 2, and 3 space dimensions. Although the effective potential is by definition time-independent, we already find a very interesting structure for the static ground states of the system. There is a close analogy between the statics of the massless KPZ system and the static behaviour of massless $\lambda \phi^4$ QFT---and much of the vacuum structure of the massless $\lambda \phi^4$ QFT carries over into the ground state structure of the massless KPZ equation. It is important to underscore the fact that the analysis presented here has focussed on the short distance or ultraviolet properties of the KPZ equation, which as we have seen, reveal a complementary phenomenology to the perhaps more common studies concerned with the infrared, or long distance properties of the KPZ equation. This distinction shows up, among other places, in the requirement of the bare tadpole term, the structure of the ultraviolet divergences requiring renormalization, the phenomena of dynamical symmetry breaking, and the marked dimension dependence of the one-loop ultraviolet renormalization group equation, and its associated fixed point. The generality of these structural and dimension-dependent features are confirmed by studying the (one-loop) effective action associated to the KPZ equation; these general results will be presented elsewhere. In 1 and 2 space dimensions we have exhibited the occurrence of dynamical symmetry breaking. In the hydrodynamic interpretation symmetry breaking corresponds to instability of the zero-velocity background, leading to the onset of bulk flows in the fluid. In the surface growth interpretation symmetry breaking corresponds to instability of the planar interface, leading to a domain structure wherein different domains exhibit different slopes (all of the same magnitude). Thus even the static ground state structure of the KPZ equation is surprisingly rich, considerably richer than one might have reasonably expected. These considerations lead one to conjecture that the one-loop methods developed in \cite{HMPV-spde} and illustrated here, may be profitably applied to reveal the onset of instabilities in the Kuramoto--Sivashinki (KS) equation. The KS equation, used to model flame-front propagation, may be regarded as a generalization of the KPZ equation with an additional fourth-derivative term $(\vec \nabla^4 \phi)$. In two dimensional combustion, recent experiments have demonstrated the existence of so-called ``fingering'' instabilities as a function of oxygen flow across the surface \cite{Fingering}. On the other hand, there is no fingering in three-dimensional combustion (because of convection). The KS equation is also Galilean invariant. Thus, it may well exhibit a dimension dependent one-loop DSB akin to that of the KPZ equation. \section*{Acknowledgments} In Spain, this work was supported by the Spanish Ministry of Education and Culture and the Spanish Ministry of Defense (DH and JPM). In the USA, support was provided by the US Department of Energy (CMP and MV). The research of CMP is supported in part by the Department of Energy under contract W-7405-ENG-36. Additionally, MV wishes to acknowledge support from the Spanish Ministry of Education and Culture through the Sabbatical Program, to recognize the kind hospitality provided by LAEFF (Laboratorio de Astrof\'\i{}sica\ Espacial y F\'\i{}sica\ Fundamental; Madrid, Spain), and to thank Victoria University (Te Whare Wananga o te Upoko o te Ika a Maui; Wellington, New Zealand) for hospitality during final stages of this work.
\section{Introduction} \label{sec:Introduction} Cir~X-1 is one of the most puzzling X-ray binaries known. Like the peculiar systems SS 433 and Cyg X-3, it cannot easily be classified into any of the major categories of X-ray binaries. Indeed, there is even doubt as to whether it is a high-mass (HMXB) or low-mass X-ray binary (LMXB). Since its discovery in early 1970s, Cir~X-1 has been studied intensively at X-ray wavelengths. The X-ray properties of Cir~X-1 were found to differ dramatically each time it was observed (see for example Kaluzienski et~al. 1976, Tennant 1988, Tsunemi et~al. 1989, Shirey {et~al.}\ 1996).\nocite{khbs76,ten88,tkm+89,sblm96} Periodic modulation of the X-ray flux was found at a period of 16.6~d \cite{khbs76}. A radio counterpart was found \cite{cpc75}, and was found to flare at the same period as the X-ray modulation \cite{hjm+78}. These flares were initially detected at peak flux levels of $> 1\;$Jy; since the 1970s, the flux of the source has decreased dramatically, and it has only occasionally been detected above $50\;$mJy \cite{snp+91}. This radio source is located $25'$\ from the centre of the supernova remnant G321.9$-$0.3, and is apparently connected to the remnant by a radio nebula \cite{hkl+86}. Stewart {et~al.}\ \shortcite{schn93} have imaged arcmin-scale collimated structures within the surrounding nebula, suggesting an outflow from the X-ray binary. Fender {et~al.}\ \shortcite{fst+98} have imaged an arcsec-scale asymmetric jet aligned with these larger structures, raising the possibility that the outflow from the system is relativistic. Recently, Case \& Bhattacharya \shortcite{cb98} have revised the estimated distance to G321.9$-$0.3 (and hence to Cir~X-1, assuming they are associated) to $5.5\;$kpc, which is substantially smaller than the original suggested distance to Cir~X-1 of $10\;$kpc \cite{gm77}. The discovery of Type I X-ray bursts \cite{tfs86b} suggests that the compact object is probably a weakly-magnetised neutron star. The close association of Cir~X-1 with the supernova remnant suggests that the system may be a young ($< 10^5$~y old) runaway system from a supernova explosion \cite{schn93}. The optical counterpart to Cir~X-1 was identified as a highly-reddened star with strong H$\alpha$\ emission \cite{wmw+77}. This object was later shown to consist of three stars within a radius of 1\farcs5, the southernmost of which is the true counterpart (\cite{mon92,dsh93}). \begin{figure*} \centerline{\psfig{figure=jfw_fig1.ps,width=\textwidth,clip=t}} \caption{Spectrum of Cir X-1, taken near apastron on 1997 June 4: the spectrum shown is the sum of five 1800~s exposures. Emission lines of lines of H$\alpha$\ $\lambda$6563, He$\,${\sc i}\ $\lambda$6678, and He$\,${\sc i}\ $\lambda$7065 can be seen; the dip at $\lambda$6870 is due to imperfect removal of the terrestrial atmospheric B-band.}\label{fig:spectrum} \end{figure*} The long orbital period and periodic X-ray activity suggested a high-mass system in an eccentric orbit \cite{mjh+80}; however, the variability of the optical emission, the faintness of the optical counterpart, and several of its X-ray characteristics suggest that the companion is a low-mass star. The lack of spectroscopic studies in the optical band means that most of the fundamental orbital parameters of Cir~X-1 have not been determined. Moreover, Cir~X-1 shows very different properties from time to time. This make it difficult to construct a coherent picture for the system from observations of different wavelengths at different epochs. Here we present new spectroscopic observations of Cir~X-1, and use these, together with analysis of archival observations of the system, to suggest a more coherent model for the system. \section{Observations and data reduction} \label{sec:Observ-data-reduct} \subsection{New optical observations} \label{sec:New-optic-observ} Cir~X-1 was observed on 1997 June 4 using the 3.9~m Anglo-Australian Telescope (AAT). The mean orbital phase of the observation was 0.51, calculated according the ephemeris of Stewart {et~al.}\ \shortcite{snp+91}. The RGO Spectrograph was used in combination with the TEK 1k CCD in the 82~cm camera and a grating of 270 grooves$\;\mathrm{mm}^{-1}$\ in first order, resulting in a dispersion of $\sim 1.08\;\mathrm{\AA}\;\mathrm{pixel}^{-1}$\ over a wavelength range 6060--7165$\;$\AA. The spatial scale was $0\farcs25$; the spectral resolution, measured from the arc lines, was $5.4\;\mathrm{\AA}$. A $1\farcs5$-wide slit was used, oriented north-south so both Cir~X-1 and star 2 of Moneti \shortcite{mon92} were in the slit. The atmospheric seeing was about $1''$. Five 1800$\;$s integrations were taken, interspersed with CuAr arc-lamp exposures, before cloud prevented the acquisition of any more data. The bias and pixel-to-pixel gain variations were removed from each exposure using standard procedures in {\sc iraf}. Cosmic rays were removed using the method of Croke \shortcite{cro95} to compare adjacent frames. Because of the presence of the nearby confusing star (Moneti's star 2), special care needed to be taken to measure the flux from our object. At every position along the dispersion direction, we fit two gaussian profiles, with fixed widths (FWHM=5.6~pixels) and separation (6 pixels), to the sky-subtracted frames. The amplitude of these gaussians was used as the estimate of the flux from Cir~X-1 and star~2 at each wavelength. We then determined the wavelength calibration using the CuAr arc lamp exposures. We fit a low-order polynomial to the arc line wavelengths as a function of pixel number: the rms scatter of the fits was $\sim 1/4$\ of a pixel. A rough flux-calibration was performed by comparing with the spectrum of the observed flux standard LTT~4364, though since the night was non-photometric, this flux calibration should be considered only approximate. \subsection{Infrared observations} \label{sec:Infr-observ} $K$-band spectroscopy of Cir~X-1 was obtained using the Cryogenic Array Spectrometer/Imager (CASPIR) on the ANU 2.3~m telescope at Siding Spring Observatory on the night of 1997 June 20. The $K$\ grism was used with the SBRC $256\times256$\ InSb array, giving a dispersion of $21\farcs5$~\AA$\,{\mathrm pixel}^{-1}$\ over a wavelength range of 1.94--2.49\ifmmode{\;\mu{\rm m}}\else$\;{\mu{\rm m}}$\fi. The spatial scale was $0\farcs5\;{\mathrm pixel}^{-1}$. A $5''$\ slit was used, oriented east-west: note that this means that Moneti's stars 2 and 3 both contributed light in our spectrum. The telescope was nodded by $\pm\, 12''$\ along the slit to provide sky frames at the same position as the object. Argon lamp spectra were taken to perform wavelength calibration, and two nearby bright stars (BS~5699 and BS 5712) were observed in order to remove atmospheric spectral features and perform flux calibration. Standard data reduction procedures were followed, using the local {\tt caspir} package running in {\sc iraf}. Bias and dark frames were used to linearise all frames, the sky background was subtracted from the object frames, and pixel-to-pixel variations were corrected. The chip distortion was corrected in order to align the dispersion and spatial directions along rows and columns of the chip; the sky background was then subtracted and spectra extracted. A low-order polynomial was fit to the argon lines, and these calibrations applied to the object spectra. Flux calibration was achieved by dividing the observed spectra by the spectrum of a nearby mid G-type star and then multiplying by a model for the absolute flux distribution of the calibrator. Residual terrestrial atmospheric features were then corrected using an early-type star. \section{Results} \label{sec:Results} \subsection{Optical spectra} \label{sec:Optical-spectra} The optical spectrum is shown in Fig.~\ref{fig:spectrum}. The spectrum clearly shows three narrow emission lines --- H$\alpha$\ $\lambda$6563, He$\,${\sc i}\ $\lambda$6678, and He$\,${\sc i}\ $\lambda$7065 --- as well as a broad component to the H$\alpha$\ line which is blue-shifted with respect to the narrow component. We fit gaussian models to these lines, using the {\tt specfit} package in {\sc iraf}. After normalising the spectrum by a low-order polynomial fit to the continuum, we fit a single gaussian to each of the helium lines and two gaussians to the H$\alpha$\ line. The velocities and widths of the three narrow lines are consistent with being the same; we therefore constrained them to be equal. The details of the fit are shown in Table~\ref{tab:optspec}. The narrow lines show a velocity of $\sim +378\kms$\ and a width of $\sim 9.5\;$\AA, or $400\kms$, while the broad component to the H$\alpha$\ line has a very different velocity ($-310\kms$, or $-688\kms$\ with respect to the narrow lines) and width ($46\;$\AA, or $2200\kms$). \begin{table} \caption{Fit to the optical spectrum. The spectrum was normalised by a low-order polynomial fit to the continuum, and four gaussians were fit to the lines, subject to the constraint that the velocity and width of the three narrow lines were the same. The table shows the resultant velocities, FWHMs, equivalent widths $W_\lambda$\ of the gaussians, and the de-reddened line luminosities, assuming a reddening of $A_V = 11$\ (Predehl \& Schmitt 1995) and a distance of $5.5\;$kpc (Case \& Bhattacharya 1998).}\label{tab:optspec}\addtolength{\tabcolsep}{-1pt} \begin{tabular}{l r@{\,$\pm$\,}l r@{\,$\pm$\,}l r@{\,$\pm$\,}l c} Component & \multicolumn{2}{c}{Velocity} & \multicolumn{2}{c}{FWHM} & \multicolumn{2}{c}{$W_\lambda$} & $L_{\rm line}$ \\ & \multicolumn{2}{c}{($\mathrm km\,s^{-1}$)} & \multicolumn{2}{c}{($\mathrm km\,s^{-1}$)} & \multicolumn{2}{c}{(\AA)} & (${\rm erg\,s^{-1}}$) \\[10pt] H$\alpha$ broad & $-$310 & 40 & 2210 & 74 & 50 & 2 & \raisebox{-1.5ex}{6.2\ee{35}} \\ H$\alpha$ narrow & +378 & 12 & 402 & 8 & 74 & 1 & \\ He$\,${\sc i}\ $\lambda$6678 & \multicolumn{2}{c}{''} & \multicolumn{2}{c}{''} & 3.4 & 0.5 & 1.5\ee{34} \\ He$\,${\sc i}\ $\lambda$7065 & \multicolumn{2}{c}{''} & \multicolumn{2}{c}{''} & 4.9 & 0.5 & 2.1\ee{34} \\ \end{tabular} \end{table}\nocite{ps95}\nocite{cb98} No stellar features can be seen in the spectrum. In particular, no absorption features can be discerned which might sugggest the nature of the binary companion. There is no significant variability apparent between the five spectra. Mignani, Caraveo \& Bignami (1997)\nocite{mcb97} obtained a spectrum of Cir~X-1 using the HST {\em Faint Object Spectrograph} near periastron (phase 0.18) in June 1995. The line profile of the H$\alpha$\ line in their spectrum was very different to our 1997 spectrum. They found the broad component of the gaussian to be centered at approximately its rest wavelength, while the narrow component was observed at a velocity of +380\kms. They interpreted the narrow component as arising in an accretion disc, with the $\sim 400\kms$\ velocity representing the rotation velocity of the disc material; the absence of the blue component they interpreted as a phase-dependent shadowing effect. Under this model, the velocity of the narrow component should shift with orbital phase, while the broad component remains fixed. Our spectrum was taken at apastron, and does not show the predicted shift of the narrow component. In fact, the velocities we observe for the narrow component of the H$\alpha$\ line (and the helium lines, not observed in the HST spectrum) are identical with the HST spectrum, while the {\em broad} component has shifted in velocity. This would seem to indicate that the model proposed by Mignani et al. for the system is wrong. \begin{table*} \begin{minipage}{\textwidth} \caption{List of observations of Cir X-1, including the archival AAT observations, the HST observation (June 1995), and our new AAT observation (June 1997). Column 2 gives the instrument/detector combination where `B\&C' indicates the Boller \& Chivens spectrograph, `RGO' indicates the RGO spectrograph with either the long (82~cm) or short (25~cm) camera; columns 3--5 show the instrumental resolution, total exposure time, and orbital phase of the observation. The following columns show results of the fit to the H$\alpha$\ line (see text for details), with gaussian 1 being the red-shifted (narrow) component, and gaussian 2 the blue-shifted (broad) component. The phase was calculated according the ephemeris of Stewart et al. (1991), and the fits are shown in Fig.~2.}\label{tab:archival-obs} \addtolength{\tabcolsep}{-1.5pt} \begin{tabular}{llccc r@{$\,\pm\,$}l r@{$\,\pm\,$}l r@{$\,\pm\,$}l r@{$\,\pm\,$}l r@{$\,\pm\,$}l r@{$\,\pm\,$}l} & & & & & \multicolumn{6}{c}{\hrulefill~gaussian 1~\hrulefill} & \multicolumn{6}{c}{\hrulefill~gaussian 2~\hrulefill} \\ UT Date & Instrument & $\Delta\lambda$ & $t_{\mathrm exp}$ & Phase & \multicolumn{2}{c}{velocity} & \multicolumn{2}{c}{FWHM} & \multicolumn{2}{c}{$W_\lambda$} & \multicolumn{2}{c}{velocity} & \multicolumn{2}{c}{FWHM} & \multicolumn{2}{c}{$W_\lambda$} \\ & & (\AA) & (s) & & \multicolumn{2}{c}{(${\rm km\,s^{-1}}$)} & \multicolumn{2}{c}{(${\rm km\,s^{-1}}$)} & \multicolumn{2}{c}{(\AA)} & \multicolumn{2}{c}{(${\rm km\,s^{-1}}$)} & \multicolumn{2}{c}{(${\rm km\,s^{-1}}$)} & \multicolumn{2}{c}{(\AA)} \\[10pt] 1976 May 21 & B \& C/IDS & 15 & 3840 & 0.536 & 370 & 40 & 1050 & 40 & 380 & 40 & 200 & 60 & 2760 & 340 & 200 & 30 \\ 1977 Feb 5 & RGO25/IPCS & 0.8 & 800 & 0.234 & 270 & 20 & 890 & 70 & 214 & 13 & $-$830 & 120 & 1200 & 160 & 57 & 10 \\ 1977 Sep 4 & RGO25/IPCS & 4.3 & 2000 & 0.947 & 490 & 50 & 640 & 50 & 245 & 25 & 160 & 120 & 2080 & 190 & 180 & 25 \\ 1978 Apr 29--30 & RGO82/IPCS & 0.45 & 10900 & 0.315 & 410 & 40 & 720 & 55 & 52 & 6 & $-$190 & 70 & 630 & 80 & 21 & 6 \\ 1978 Aug 11 & RGO25/IPCS & 1.3 & 2000 & 0.522 & 220 & 200 & 700 & 600 & 180 & 130 & $-$600 &2000 & 1030 & 120 & 30 & 100 \\ 1995 Jun 1 & HST/FOS & 2.2 & 5160 & 0.183 & 330 & 10 & 355 & 15 & 30 & 2 & $-$50 & 20 & 1145 & 25 & 75 & 2 \\ 1997 Jun 4 & RGO82/TEK & 5.4 & 9000 & 0.510 & 380 & 12 & 400 & 10 & 74 & 2 & $-$300 & 50 & 2210 & 100 & 50 & 2 \\ \end{tabular} \end{minipage} \end{table*} \begin{figure} \centerline{\psfig{figure=jfw_fig2.ps,width=6.6cm,clip=t}} \caption{Line profiles of H$\alpha$\ in the archival observations of Cir X-1, showing the two gaussians fit to each line and their sum. Each spectrum had been binned to twice the spectral resolution, and normalised by a polynomial fit to the continuum, so the $y$-axis shows the relative flux of the line. The last two spectra show the HST observation (June 95) and our new AAT observation (June 97).}\label{fig:archivalfit} \end{figure} To investigate this, we used data from the AAT archive, which contains nearly all observations taken with the AAT since its inception. Cir~X-1 has been observed using the AAT several times by different observers, with most observations occurring about 20 years ago. We obtained spectra at a total of five epochs which showed detectable emission: these are listed in Table~\ref{tab:archival-obs} and shown in Fig.~\ref{fig:archivalfit}. The May 1976 data were published by Whelan {et~al.}\ \shortcite{wmw+77}; the rest of the data are unpublished. We obtained the original data from the AAT archive and reduced them using standard techniques and the {\sc figaro} data reduction package. Note that for all these observations, light from stars 2 and 3 was presumably in the slit. Because of this, no attempt has been made to flux calibrate the spectra; instead, each spectrum was binned to a spectral sampling of two data points per resolution element, and the resulting spectrum was normalised by a low-order polynomial fit to the spectrum. We performed the same double gaussian fit to these spectra as described in Table~\ref{tab:optspec}, and the results are shown in Table~\ref{tab:archival-obs}. Several points are immediately clear: despite the wide range of orbital phases, the broad component is never observed redward of the narrow component\footnote{Except for the Apr 78 spectrum, where the width of the red-shifted component is marginally smaller than the width of the blue-shifted component; however, the two widths are consistent with being identical.}. The (weighted) mean velocity of the narrow component is $+350\kms$, that of the broad component $-100\kms$. The scatter of the velocities of the broad component is much greater than that of the narrow component. No strong trends can be seen with either date of observation or orbital phase, with the exception of the equivalent width of the narrow component, which has been generally decreasing since 1976 (except for the extremely low value in April 1978). \subsection{Infrared spectra} \label{sec:Infrared-spectra} Infrared spectra were obtained one orbit after the optical spectra. The mean spectrum is shown in Fig.~\ref{fig:IRspectrum}. Bright emission lines of Br~$\gamma$\ and He$\,${\sc i}\ $\lambda$2.058\ifmmode{\;\mu{\rm m}}\else$\;{\mu{\rm m}}$\fi\ were visible in the spectrum. These two lines are generally observed in the K-band spectra of low- and high-mass X-ray binaries (\cite{bsc+97,bsct98,csfm99}) and may arise in accretion discs or larger emitting regions, such as dense winds from the mass donors and/or ejecta from the binaries. There also seems to be a hint of asymmetry in the blue wing of the infrared lines, so we again fit gaussians to the line profile. We modelled each emission line as the sum of two gaussians, constraining the widths and velocities of the blue and red components to be the same for both lines. The results are shown in Table~\ref{tab:IRspec}. \begin{figure} \centerline{\psfig{figure=jfw_fig3.ps,width=\columnwidth,clip=t}} \caption{Infrared spectrum of Cir X-1, taken near apastron on 1997 June 20.}\label{fig:IRspectrum} \end{figure} Both lines are well fit by one gaussian at $+450\kms$, a velocity very similar to the narrow red-shifted optical component, plus a blue-shifted component at a velocity of $\sim -2000\kms$. \begin{table} \caption{Fit to the infrared spectrum. The spectrum was normalised by a low-order polynomial fit to the continuum, and four gaussians were fit to the lines, subject to the constraint that the velocity and widths of the blue (B) and red (R) components were the same for both species. The final column shows the de-reddened integrated line luminosities, assuming a reddening of $A_V = 11$\ and a distance of $5.5\;$kpc.}\label{tab:IRspec}\addtolength{\tabcolsep}{-1pt} \begin{tabular}{l r@{\,$\pm$\,}l r@{\,$\pm$\,}l r@{\,$\pm$\,}l c} Component & \multicolumn{2}{c}{Velocity} & \multicolumn{2}{c}{FWHM} & \multicolumn{2}{c}{$W_\lambda$} & $L_{\rm line}$ \\ & \multicolumn{2}{c}{(\kms)} & \multicolumn{2}{c}{(\kms)} & \multicolumn{2}{c}{($\mu{\rm m}$)} & (${\rm erg\,s^{-1}}$) \\[10pt] H$\,${\sc i}\ Br$\gamma$\ (R) & 450 & 40 & 1170 & 90 & 88 & 7 & 4.3\ee{34} \\ He$\,${\sc i}\ (R)& \multicolumn{2}{c}{''} & \multicolumn{2}{c}{''} & 96 & 7 & 3.1\ee{34} \\ H$\,${\sc i}\ Br$\gamma$\ (B) & $-$1870 & 570 & 1020 & 440 & 12 & 6 \\ He$\,${\sc i}\ (B)& \multicolumn{2}{c}{''} & \multicolumn{2}{c}{''} & 11 & 5 \\ \end{tabular} \end{table} \section{An eccentric low-mass binary model for Cir X-1} \label{sec:new-model} Initially, it was suggested that Cir~X-1 is a high mass binary consisting of a compact star (either a neutron star or black hole) and an OB supergiant companion star (\cite{wmw+77,mjh+80}). Its X-ray properties imply a very eccentric binary orbit. The 16.6-day modulation in the X-ray luminosity is due to orbital variations in the mass accretion of the compact star. The X-ray and radio bursts occur when the compact star encounters the dense stellar wind from the supergiant \cite{ff80}, or when tidal mass transfer is induced during the periastron passage \cite{hay87}. The Type I X-ray bursts that were discovered in a brief epoch in the mid-1980s \cite{tfs86} indicated that the compact star is a neutron star. Recent observations ({e.g.}\ \cite{snp+91,gla94}) indicate that the companion star is unlikely to be a massive supergiant star. No Type I bursts have yet been reported in RXTE observations. Here, we propose a low-mass binary model, where the system consists of a neutron star orbiting around a subgiant companion star of about 3 to 5\Msolar\ (Fig.~\ref{fig:model}; see also Tennant \& Wu 1998).\nocite{tw98} The orbital eccentricity is $\sim 0.7$--0.9. During the periastron passage, the companion star overfills its Roche-lobe, causing a transfer of mass at a super-Eddington rate onto the neutron star. As Cir X-1 is a X-ray burster, the magnetic field of the neutron star is relatively weak, and so the accretion flow is probably quasi-spherical during the periastron passage. Because of the large radiative pressure of emission from the neutron star, there must also be a strong (anisotropic) matter outflow. The inflow/outflow geometry may be related to the larger scale jets observed from the system, with symmetrical collimated outflows along some preferred axis. \begin{figure} \centerline{\psfig{figure=jfw_fig4.ps,width=7cm,clip=t}} \caption{Model for Cir X-1. During periastron (phase 0) the neutron star passes its closest to the companion (mass $\sim 3$--$5\protect{\Msolar}$), and the disc is disrupted, resulting in large X-ray fluctuations. After periastron passage, an accretion disc begins to form; mass accretion steadies, until after apastron (phase 0.5) steady accretion via a disc takes place. Tidal interactions begin to disrupt the disc again as periastron approaches.}\label{fig:model} \end{figure} After the periastron passage, the companion star is detached from its critical Roche-surface, and mass transfer ceases. Accretion continues as the neutron star captures the residual matter in its Roche-lobe. An accretion disc is gradually formed. The disc is geometrically thick, because the accretion rate is sufficiently high and is close to the Eddington limit. The accretion becomes more steady when disc accretion takes over from the less stable super-Eddington quasi-spherical accretion. Accretion via a disc continues to operate till the next periastron passage, where the accretion disc is disrupted by tidal interaction. When the periastron passage proceeds, the companion star overfills its Roche-lobe again, leading to another cycle of super-Eddington mass transfer, formation of the accretion disc, steady accretion via the disc and disruption of the accretion disc. The periodic X-ray light curve of Cir X-1 has changed dramatically since its discovery in the 70s ({e.g.}\ \cite{khbs76,tkm+89,sblm96}). This indicates that the accretion mode has varied substantially in the last 25 years. Despite the fact that the H$\alpha$\ lines in the optical spectra obtained in the 70s and 90s show certain similarities, such as an asymmetric profile with a broad blue wing, it is inappropriate to assume that the optical spectra are the same at the same orbital phases in the 70s and the 90s. The FWHM of the red (narrow) component of the H$\alpha$\ line seems to decrease from $\approx 700\kms$\ in the 70s to $\approx 380\kms$\ in the 90s. The central wavelength, however, does not show significant velocity variation, and its average value is $350\,\pm\,80\kms$. Our interpretation is that the red (narrow) component is emitted from irradiatively-heated matter which is relatively stationary with respect to the centre of mass of the system and has a low velocity dispersion. One of the natural choices is the heated surface of the companion star, on which the projected velocity of the matter is much less than the flow velocity near the neutron star.\footnote{Note that, if the narrow emission is arising from the heated face of the secondary, the equivalent width should vary in a smooth fashion through the binary orbit, with a maximum at phase 0.5; such a pattern is not seen in the data in Table~\ref{tab:archival-obs}. However, we argue that the mass transfer rate has varied so much over the 21~y spanned by the observations that it is misleading to try to compare the equivalent widths. Indeed, if we consider three groups of spectra near to each other in time (1976--1977, 1978, and 1995--1997) there is a (slight) tendency for the equivalent width to be highest at phase 0.5. Such a relation certainly needs to be confirmed.} The velocity of the sharp red component of the H$\,${\sc i}\ Br$\gamma$\ and the He$\,${\sc i}\ $\lambda 2.058\ifmmode{\;\mu{\rm m}}\else$\;{\mu{\rm m}}$\fi$\ lines is $450\,\pm\, 40\kms$, which is consistent with that for the optical lines. Thus, the lines must also be emitted from a region relatively stationary to the centre-of-mass of the binary. The width of the IR lines, however, is much larger than the width of the optical lines. It is therefore unlikely that the emission site is the surface of the companion. One possibility is that these IR lines are emitted from a heated dust shell -- the residue of ejecta from previous epochs. The observations accumulated since the 70s show that the broad component is blueward of the narrow component, except (possibly) in the 1978 April observation. The line widths were generally above 1000\kms, and it was even larger than 2000\kms\ in the 1976 and 1997 observations. Although the line centre velocities appear to show large variations (from $-600\kms$\ to $200\kms$), they show no obvious correlation with orbital phase. We suspect that the broad component is related to the high velocity flow from regions near the neutron star. X-ray observations have shown that the rate of mass accretion onto the neutron star is above the Eddington limit (e.g.\ Inoue 1989).\nocite{ino89} It is therefore possible that an outflow is driven by the Eddington luminosity. As we only see the blueward component, the lines are probably emitted from the optically thick outflowing matter, of which the red-shifted components are heavily absorbed or obscured. The changes in the line centre velocities and widths deduced from the fits are probably the result of the changes in optical depth over the different observational epochs. \begin{figure} \centerline{\psfig{figure=jfw_fig5.ps,width=\columnwidth,clip=t}} \caption{Line profile of emission from an expanding envelope, with parameters as described in the text. Positive x corresponds to the blue-shifted frequency; negative x corresponds to the red-shifted frequency. The filled circles are the simulated data. The solid line is a fitted model with a constant and two gaussians. The two gaussians are represented by the dotted lines. The ratio of the widths of the broad and narrow gaussians is 5.5, and the ratio of the width of the broad gaussian to the separation of the two gaussian peaks is 3.22.}\label{fig:sim-main} \end{figure} The quality of our data precludes us performing a fit of this model to the line profile; rather, we present it as a possible explanation for the observed asymmetric line profile. We have constructed a simple toy model to demonstrate the asymmetric line profile that might be observed from such an outflow. In our model the broad line emission region is a spherically symmetric extended envelope, with outer and inner radii $R_{\rm out}$\ and $R_{\rm in}$\ respectively. The envelope encloses an opaque sphere with radius equal to $R_{\rm in}$. We have considered various functional forms for the velocity, emissivity and absorption coefficient distribution, and have found that models with a power-law velocity profile generally can produce a broad line profile similar to the observations (see Appendix). In Fig.~\ref{fig:sim-main}, we show a simulated a blue-shifted broad line from an expanding envelope with outer and inner radii, $R_{\rm out} = 1$\ and $R_{\rm in} = 0.3$\ respectively. The radial expanding speed $V$\ has a power-law profile with an index of +1, and it is normalised such that its value is 1 at $R_{\rm out}$. The normalised local Doppler width of the line is 0.1. The emissivity is uniform within the emitting region, with $\eta_c = \eta_l =1$. The continuum and line absorption coefficients are uniform, and their ratio $\chi_l/\chi_c = 0.35$. The total effective absorption optical depth $\tau$\ is such that $(\chi_c + \chi_l) R_{\rm out} = 1$. As the narrow line is emitted from the heated companion star, we simply assume it is a gaussian with a width of 0.123 and a norm of 4.0. This will give an approximate ratio of 2/3 for the equivalent widths of the simulated broad and narrow line components. As shown, the resultant line profiles resembles the H$\alpha$\ line observed in our June 1997 data. If we fit the simulated line with two gaussians and a constant, the relative widths and shifts of the two gaussian components are in good agreement with the fitted parameters of the June 1997 data (Fig.~\ref{fig:sim-main}). The blue components of the H$\,${\sc i}\ Br$\gamma$\ and the He$\,${\sc i}\ $\lambda 2.058 \ifmmode{\;\mu{\rm m}}\else$\;{\mu{\rm m}}$\fi$\ lines have a velocity shift of $-1870\,\pm\, 570\kms$\ and a width of $1020\,\pm\, 440\kms$. If we interpret the blue component of the IR lines as being emitted from the vicinity of the neutron star, the difference in the velocities between the optical and IR lines is due to optical depth effects and/or the uncertainties in determining the central velocities of the lines with the spectral resolution of the IR spectrum. Murdin {et~al.}\ \shortcite{mjh+80} assumed a 1\Msolar\ compact star, and by comparing the radius of various classes of stars with the separation of the two component star at periastron, they estimated the mass of the companion star and the orbital eccentricity for which tidally induced mass transfer can occur during the periastron passage. In our model, the mass transfer is driven by the overflow of the mass-donor star's Roche-lobe instead of the stellar wind from the star, and it therefore requires the companion star to fill its Roche-lobe at periastron, {i.e.}\ the star's radius $R_2$\ to be equal to or larger than its `instantaneous' critical Roche-lobe surface. It is beyond the scope of this paper to derive an exact phase-dependent effective Roche-lobe surface. However, as a first approximation, if we simply take the expression of the Roche-lobe radius for circular systems (see Kopal 1959)\nocite{kop59} and replace the orbital separation term by the separation between the two stars of Cir X-1 at periastron, then we can obtain an effective Roche-lobe radius at periastron (phase~0): \begin{equation} R_{_{\mathrm L}} \approx a(1-e) [ 0.38 + 0.2 \log q ] \end{equation} where the mass ratio $q = M_2/M_1$\ and $a$\ is the semi-major axis. Using this approximate expression for $R_{_{\mathrm L}}$, we can constrain the appropriate parameter space ($e$; $M_1$, $M_2$) where Roche-lobe overflow occurs during the periastron passage for various types of mass-donor stars. \begin{figure} \centerline{\psfig{figure=jfw_fig6.ps,width=6.5cm,clip=t}} \caption{The radius of the Roche-lobe at periastron as a function of the mass of the companion star, calculated from an eccentric system with a 16.6~d period and a 1.4\Msolar\ neutron star primary. For Roche-lobe overflow to occur, the size of the secondary must exceed or equal the volume enclosed by the instantaneous critical Roche-lobe surface. We also show the size of the star for various masses at the zero-age main-sequence (ZAMS), terminated main-sequence (TAMS), giant and supergiant stages (indicated with filled triangles, open triangles, open squares and open circle respectively). The radii of the stars are derived from the evolutionary models of Maeder \& Meynet (1988), with the ZAMS, TAMS and giant stages corresponding to Point 1, 7 and 14 in the evolutionary tracks. The mass-radius relation of the super-giant stars, which is not as well-defined, is adopted from Lang (1991).}\label{fig:RLradius} \end{figure}\nocite{lan91}\nocite{mm88} In Fig.~\ref{fig:RLradius} we show the mass-radius relation for supergiant, giant and main-sequence stars, together with $R_{_{\mathrm L}}$\ for eccentric systems with a 16.6 day period and a 1.4\Msolar\ neutron star primary. If the companion star is a main sequence star, its mass is about 10\Msolar\ and the orbital eccentricity $e > 0.9$\ in order to fill its Roche-lobe at periastron. A sub-giant star with $M_2 \approx 3$--$5\Msolar$\ can easily fill its Roche-lobe for an orbital eccentricity of about 0.7--0.9. The argument for a low mass system is generally consistent with the combined results of our spectroscopic observations and the study on the effects of supernova kicks on neutron star binaries. A simulation study \cite{bp95b} showed that the eccentricity of HMXBs (with $M_2 \sim 10\Msolar$) and LMXBs (with $M_2 \sim 1$--$5\Msolar$) resulting from supernova explosions with isotropic kicks will be about 0.5 ($\sigma \approx 0.2$) and 0.8 ($\sigma \approx 0.05$) respectively if the final orbital period is about 16 days. The corresponding centre-of-mass velocities are 50 ($\sigma \approx 12$) and 150 ($\sigma \approx 100$)\kms. If we accept the interpretation that the narrow line components originate from the heated companion star, the line-of-sight component of the centre-of-mass velocity of the system is about 350\kms, or 430\kms after applying a correction for Galactic rotation (using the rotation curve of Clemens 1985)\nocite{cle85}. This correction was made assuming that the position of G321.9$-$0.3 was the birthplace of Cir~X-1, but this makes little difference to the resulting correction. By combining with the transverse velocity ($\sim 300$--$400\kms$) inferred from the age and the distance of the supernova remnant G321.9$-$0.3 \cite{cpc75} we obtain an estimated centre-of-mass velocity of $\sim 540\kms$. This velocity could be substantially higher, as it depends on the (very uncertain) age estimate of the supernova remnant; transverse velocities as high as 1600\kms are possible \cite{schn93}. A velocity of $\sim 540\kms$\ is about 3$\sigma$\ from the mean centre-of-mass velocities obtained from the simulation study by Brandt \& Podsiadlowski \shortcite{bp95b}. Such high velocities are not at all consistent with the simulations for high mass systems: even the low velocity is about 13$\sigma$\ above the mean velocity predicted for a high-mass system. The implied velocity for Cir~X-1 makes it one of the fastest moving binaries known: radial velocities of $\sim 300\kms$\ have been measured for a handful of other LMXBs, {e.g.}\ 4U~1556$-$605 \cite{mpmb89}, GX~349+02 \cite{pa91}, 4U~1957+11 \cite{mtb78}; see Johnston \shortcite{joh92} for a discussion. Such velocities strengthen the identification of Cir~X-1 as a low-mass system, as the fastest moving HMXBs have velocities $v < 100\kms$. If the transverse velocity is similar to the radial velocity, this implies a proper motion of 15~${\mathrm mas}\,{\mathrm yr}^{-1}$; in an ongoing project to measure the proper motion of the radio counterpart to Cir~X-1, we can measure a relative positional accuracy of around 40~mas, so we should measure the proper motion in a couple of years. An alternative simulation by Shirey \shortcite{shi98} also concluded that Cir~X-1 is a low-mass system. However, his simulation requires the system to be old ($\sim 10^7$~y), which seems difficult to reconcile with the observations that Cir~X-1 is associated with the supernova remnant G321.9$-$0.3. The RXTE observations show that the current X-ray luminosity is quite steady in the orbital phases around 0.5--1.0. The cycle-to-cycle variations are insignificant in comparison with the those at the earlier orbital phases. We interpret the steady accretion as due to the presence of an accretion disc. The erratic cycle-to-cycle luminosity variations are the consequence of the absence of the disc acting as a buffer; cf. the rms fluctuations in the RXTE data at phases 0.15 and 0.5 \cite{sblm96}. Moreover, we do not see double-horn features in the H$\alpha$\ line, such as those often observed in cataclysmic variables ({e.g.}\ \cite{hm86}) or black hole binaries ({e.g.}\ \cite{swhw98}). This, together with the near-Eddington luminosity of the X-rays during the disc-accretion phase, implies that the disc is non-Keplerian and it is both geometrically and optically thick. As the matter is not tightly bound by the gravity of the neutron star, the large variation in the tidal force when the neutron star approaches periastron could easily disrupt the disc, causing a short temporal decrease in the accretion luminosity of the system. What constraints can we put on the age of the system? In our model, mass transfer at periastron does not begin until the star has evolved from the main-sequence towards the giant branch. For a 3--5\Msolar\ star, the subgiant stage may last $\sim 5\ee{5}$~y. If Cir X-1 is associated with the nearby supernova remnant G321$-$0.3, then the implied age of the system as an eccentric neutron star binary is $< 10^5$~y. The Eddington limit for the rate of mass transfer onto a neutron star is $\sim 3\ee{8}\Msolar\;{\mathrm y}^{-1}$, giving a mass transfer time scale of $3\ee{7}$~y. Since the companion star in Cir X-1 can lose mass at a rate greater than this as an average in each orbital cycle, the life-span of the system as an X-ray binary is correspondingly shorter, but still comfortably above either the evolution time scale of the companion star or the age of the supernova remnant. \section{Conclusions} \label{sec:Conclusions} We have detected strong optical and infrared emission lines in the spectrum of Cir~X-1; the H$\alpha$\ and infrared lines are asymmetric, with a narrow component at a velocity of $\sim +350\kms$\ and a broader, blue-shifted component. Archival optical observations show that an asymmetric H$\alpha$\ emission line has been in evidence for the past twenty years, although the shape of the line has changed significantly. The narrow component is always seen redwards of the broad component, at a velocity around 200--$400\kms$; the broad component has a much larger scatter of velocities. These properties are not consistent with the interpretation of Mignani {et~al.}\ \shortcite{mcb97}, who suggested that the narrow component arises in an accretion disc, with the absence of a red component due to a phase-dependent shadowing effect. Instead, we suggest that the narrow component arises from the heated surface of the companion star, which is a subgiant of about 3--5\Msolar, while the broad component arises in an optically thick, high velocity outflow driven by super-Eddington accretion onto the neutron star. During periastron passage, the companion star overfills its Roche-lobe, causing a transfer of mass at a super-Eddington rate, which in turn drives a strong matter outflow. After periastron, mass transfer from the companion ceases, but accretion continues at a near-Eddington rate as the neutron star captures the residual matter in its Roche-lobe. An accretion disc gradually forms, which is non-Keplerian and geometrically thick. This change between quasi-spherical and disc accretion causes the change between strong X-ray variability near phase 0 and steady accretion between phases 0.5--1.0. Thus our observations are in general consistent with a low-mass binary model for the system. In this model, the velocity of the narrow component reflects the space velocity of the binary; the implied radial velocity (+430\kms) makes Cir~X-1 one of the fastest moving binaries known. \section*{Acknowledgments} \label{sec:Acknowledgments} We thank Allyn Tennant, Peter Wood, Lex Kaper, and Thomas Tauris for useful discussions; we also thank Peter for assistance with our IR observation. RPF was supported during the period of this research initially by ASTRON grant 781-76-017, and subsequently by EC Marie Curie Fellowship ERBFMBICT 972436. KW is supported by ARC through an Australian Research Fellowship. This research was based in part on observations made with the NASA/ESA Hubble Space Telescope, obtained from the data archive at the Space Telescope Science Institute. STScI is operated by the Association of Universities for Research in Astronomy, Inc. under NASA contract NAS 5-26555.
\section[1]{Introduction} Integrable differential equations are those that are solvable (for a large space of initial data) through an associated linear problem\cite{ac:cup}. It is conjectured that the solutions of all such equations possess a characteristic complex singularity structure \cite{ars:jmp}. In particular, there is widespread evidence that all movable singularities of all solutions are poles \cite{kjh:96,mdkpc:pp}. This is commonly referred to as the \textit{Painlev\'e\ property}. Extensions of this definition can be found for example in \cite{cfp:neg}. This property has been used as a starting point for deducing other remarkable properties of integrable partial differential equations (PDEs). The basic idea, first proposed by Weiss\cite{weiss:II} (see also \cite{wtc:jmp}), is to truncate a Laurent expansion of a generic solution near a movable pole. For example, the Korteweg-deVries (KdV) equation \[ u_t+6uu_x+u_{xxx}=0,\quad u=u(x,t) \] admits locally convergent Laurent series of the form \cite{njjap:nonlin,njgs:nonlin} \begin{equation} u = {-\,2{\Phi_x}^2\over {\Phi}^2}+{2{\Phi_{xx}}\over {\Phi}}+ \sum_{n=0}^{\infty}u_n(x,t){\Phi(x,t)}^n, \label{laurent_kdv} \end{equation} as solutions in a neighbourhood of any analytic, noncharacteristic variety given by $\Phi(x,t)=0$. (Note that noncharacteristic here implies $\Phi_x\not=0$.) The \lq\lq truncation\rq\rq\ of this series is \begin{equation} u = {-\,2{\Phi_x}^2\over {\Phi}^2}+{2{\Phi_{xx}}\over {\Phi}} +g(x,t) \label{trunc_kdv} \end{equation} where $g(x,t)$ (often called the \lq\lq constant-level\rq\rq\ term) is analytic near $\Phi=0$. Asking that this expression be a solution of the KdV equation, order by order in $\Phi$, requires that $g$ must be a solution of the KdV equation, and also that $\Phi$ satisfies an equation, often referred to as the ``singular manifold equation.'' Weiss showed how it is then possible to deduce the well known linear problem and Darboux transformation for the KdV equation. These then yield the B\"acklund\ transformation for the KdV equation. Such ideas have led to a procedure called the \lq\lq truncation method\rq\rq\ that has been successfully extended \cite{egmr93,pilar1,mc:1,cmp:2,p:96,pilar2} to many PDEs. However, no such general procedure exists for ODEs. The PDE-truncation procedure relies on setting coefficients of different powers of $\Phi$ to zero in the image equation obtained by substituting the truncation. This poses a difficulty for ODEs. In each case for which this has been tried for an ODE \cite{w:84a,w:84b,ntz:87,gntz:88}, the results have been found to be very restricted. For example, only special transformations or special solutions have been found via this procedure. In particular, no general, parameter-dependent B\"acklund\ transformation has been found by a truncation method. Recently, Clarkson, Joshi and Pickering \cite{cjp} have shown how this difficulty can be overcome for the second Painlev\'e\ (\PII ) hierarchy. The main idea here was to use the truncation and the image equation (obtained by substituting the truncation) to eliminate $\Phi$, instead of separating powers of $\Phi$. The result is a B\"acklund\ transformation for the hierarchy. This approach has also been extended to the reductions of the modified Sawada-Kotera/Kaup-Kupershmidt (mSK/mKK) hierarchy \cite{jp}. Our purpose here is to give a {\em universal} truncation-type method for ODEs that is based on singularity analysis. The above applications (to the \PII\ and reduced mSK/mKK hierarchies) can be recast in this framework. The main new idea is to consider truncation as a mapping that preserves the locations of a natural {\em subset} of movable singularities. An example is given by the second Painlev\'e\ equation \begin{equation} y''=2y^3+xy+\alpha,\quad y=y(x), \label{p2} \end{equation} whose general solution possesses two families of movable poles \cite{njmdk:direct}. (These are often referred to in the literature as two \lq\lq branches\rq\rq of a Painlev\'e\ expansion.) Near a movable singularity $x=x_0$, say, $y(x)$ has a convergent Laurent expansion \begin{equation} y = {\pm\,1\over (x-x_0)}+h(x)(x-x_0),\label{laurent_p2} \end{equation} where $h(x)$ is locally analytic. Clearly, the set of poles of $y(x)$ naturally separates into two subsets, identified by the sign of the coefficient of the leading-order pole. Each of the Painlev\'e\ equations, except the first, has generic solution $y(x)$ that possesses pairs of simple movable poles with coefficients of opposite sign. Therefore, the set of all movable poles of a solution $y(x)$ decomposes into the union of two nonintersecting subsets $\cP_+$ and $\cP_-$. By $\cP_+$ we mean the set of poles with positive choice of coefficient and $\cP_-$ is that with the negative choice. In the following section, a generic solution $y(x)$ of a Painlev\'e\ equation will be transformed to a solution $Q(x)$ of the same equation but with possibly different parameters as \begin{equation} y(x)=\rho(x) + Q(x).\label{trunc} \end{equation} We construct the transformation by demanding that $\rho(x)$ have poles exactly at the elements of $\cP_+$ and $Q(x)$ have them at $\cP_-$ (or vice-versa). When $Q(x)$ satisfies the same equation (albeit with different parameters), we find B\"acklund transformations (BTs), through a procedure that relies only on singularity analysis of the transformed equation. The method is applied to \PII\ and \PIV\ in this paper. In section 2, we give details for \PII\ and indicate the major differences for \PIV . This method also leads to related ODEs and B\"acklund\ transformations between these and the equation under consideration; these are considered in Section 3, again for \PII\ and \PIV . In section 4, we show how to carry out the analogue of the so called double singular manifold method, i.e. one which considers two singularities of the solution simultaneously. Extensions of these techniques to other Painlev\'e equations are considered in a second paper \cite{gjp2}. \section{Truncation for \PII\ and \PIV } In this section we show how to carry out a mapping that preserves \lq\lq half\rq\rq\ the movable poles of the solution of a Painlev\'e\ equation and how this can lead to a B\"acklund\ transformation for that equation. We recall the equations \PII\ and \PIV\ here for reference later. \begin{eqnarray} y''&=&2y^3+xy+\alpha\\ y''&=&{{y'}^2\over 2y}+{3 y^3\over 2}+4xy^2+2(x^2-\alpha)y-{\gamma^2\over 2y} \end{eqnarray} (Note the slightly unusual renaming of the second parameter in \PIV\ that differs from convention.) We take $Q(x)$ to satisfy the same equation as $y(x)$ but with possibly different parameters indicated by roman letters replacing the corresponding greek ones. (So e.g. $\alpha\mapsto a$.) \subsection{\PII } Substitution of Eq (\ref{trunc}) into \PII\ gives \begin{equation} \rho'' - 2\rho^3-6\rho^2 Q - 6\rho Q^2 -x\rho -\alpha +a =0. \label{rho_ode_p2} \end{equation} The dominant terms of this equation near a pole of $\rho$ are \[ \rho'' \approx 2\rho^3\ \Rightarrow\ {\rho'}^2 \approx {\rho}^4. \] Taking the positive square root (i.e. taking $\rho$ to have a pole in $\cP_-$), we write \begin{equation} \rho'(x)=:\rho(x)^2 + \sigma(x)\rho(x),\label{ricc_p2} \end{equation} where $\sigma$ is to be found. Using this to replace $\rho'$ in the equation (\ref{rho_ode_p2}), we get \[ \bigl(3\sigma(x)-6Q(x)\bigr)\rho(x)^2 +\left(-x+\sigma(x)^2+\sigma'(x) - 6 Q(x)^2\right)\rho(x)-\alpha+a=0. \] Consider the dominant terms of this equation near a pole of $\rho$. Since $Q(x)$ is regular at such points, we get to leading-order \[ \sigma(x)\approx 2Q(x)\ \Rightarrow\ \sigma(x)=:2Q(x) + {\tau(x)\over\rho(x)} \] Now Eq (\ref{rho_ode_p2}) becomes \[ \left(-x+2Q'(x)-2Q(x)^2+2\tau(x)\right)\rho(x)+ 2\tau(x)Q(x)+\tau'(x)+a-\alpha =0, \] from which we have \[ \rho(x) = -\,{2\tau(x)Q(x)+\tau'(x)+a-\alpha\over -x+2Q'(x)-2Q(x)^2+2\tau(x)}. \] However, this must be compatible with the Riccati equation (\ref{ricc_p2}) for $\rho(x)$. Substituting this expression for $\rho(x)$ into (\ref{ricc_p2}) yields a compatibility condition that contains $Q$, $Q'$ and $Q''$. We use the equation \PII\ satisfied by $Q(x)$ to eliminate $Q''$. The result is a polynomial equation in $Q$ and $Q'$, with coefficients involving $\tau$, $\tau'$ and $\tau''$, given by \begin{eqnarray} 0&=& - 4\,\tau(x)Q(x)^{4} + 4\tau'(x)Q(x)^3\nonumber \\ &&+ \left(12{\tau}(x)^{2} + 12{ \tau}(x\,)Q'(x) + 2 \tau''(x) - 4x\,\tau(x)\right) \,{ Q}(\,{x}\,)^{2} \nonumber \\ && +\left(- 2\,{\tau}(x) +2 x \tau'(x) - 4 Q'(x)\tau'(x) + 4a\,{ \tau}(x)\right)Q(x) \nonumber \\ && - 8\,{ \tau}(x) {(Q'(x))}^{2} \nonumber\\ &&+ \left( - 2{\tau''(x)} + 6x \tau(x) - 12\,{ \tau}(x)^{2}\right)Q'(x) \nonumber \\ &&- {x}^{2}{ \tau}(x) +{a}^{2} - {\alpha}^{2} + { \alpha} -a+2a \tau'(x)\nonumber\\ &&- \tau'(x) + \bigl(\tau'(x)\bigr)^{2} \nonumber \\ & &- 2\,{ \tau''(x)}\,{ \tau}(\,{x}\,) - 4\,{ \tau}(\,{x} \,)^{3} + x\,{\tau''(x)} + 4x\,{ \tau}(\,{x}\, )^{2}\label{p2_compat} \end{eqnarray} Since the solution space of \PII\ depends on two arbitrary parameters given by $Q$, $Q'$ at a (regular) point, we can demand that this equation be satisfied identically in these two variables. Since $\rho$ is a functional of $Q$ (or equivalently $y$), $\tau$ is also. The simplest solutions of Eq (\ref{p2_compat}) are those that are polynomial in $Q$ and $Q'$. Inspection of Eq (\ref{p2_compat}) shows that the simplest solution independent of both $Q$ and $Q'$ is $\tau(x)\equiv 0$ (because of the presence of the monomial $4\tau(x)Q(x)^4$) under the constraint \[ a^2-a = \alpha^2-\alpha\ \Rightarrow\ a=\alpha\ \textrm{or}\ a =-\alpha +1 . \] Then, the above solution for $\rho$ becomes \begin{equation} \rho(x) = {a-\alpha\over x-2Q'(x)+2Q(x)^2} . \label{bt1} \end{equation} Thus in addition to the identity $y(x)=Q(x)$ we obtain the well known BT for \PII\ \cite{gambier:acta}. We have also considered other possible solutions for $\tau$, e.g. those that only depend polynomially on $Q$ but not $Q'$. The monomial $- 2{\tau''(x)}Q'(x)$ in Eq (\ref{p2_compat}) shows that it can be at most linear in $Q$. Other monomials in the resulting equation then lead again to $\tau(x)\equiv 0$. (For other equations, a possible dependence of $\tau$ on $Q$ can give more general results \cite{gjp2}.) If in Eq (\ref{ricc_p2}) we had taken the negative root, i.e. \[ \rho'(x)=-\rho(x)^2 + \sigma(x)\rho(x), \] the above procedure would have yielded \[ \rho(x) = {a-\alpha\over x+2Q'(x)+2Q(x)^2}, \quad a=-\alpha-1 , \] the alternative form of the BT for \PII . This could also have been found by using the discrete symmetry of \PII\ under $y\mapsto -y$, $\alpha\mapsto -\alpha$ in combination with Eqn (\ref{bt1}). It should be noted, however, that both BTs are needed in order to iterate to find sequences of special integrals and rational solutions. \subsection{\PIV } Here we describe major differences in applying the above procedure to \PIV . The first difference lies in the dominant terms of the transformed equation. These are \[ \rho''\rho-{{\rho'}^2\over 2}\approx {3\rho^4\over 2}. \] After using the integrating factor $\rho'\rho^{-2}$, we can integrate to get \[ {\rho'^2\over \rho}\approx \rho^{3}\ \Rightarrow\ \rho'\approx\pm\rho^2. \] Consider first the case with minus sign on the right, i.e. \[ \rho'=-\rho^2+\sigma \rho. \] Then we find (following dominant balances again) that \[ \sigma(x)=-2x-2Q(x)+{\tau(x)\over\rho(x)} \] Using this in the transformed equation leads to a quadratic equation for $\rho$: \begin{eqnarray} & &\Biggl( -\, 2x Q(x) - { \tau}(x) - Q({x})^{2} - 2 + 2\,{ \alpha} - Q'({x}) \Biggr) \rho({x})^{2} +\nonumber \\ & &\quad\Biggl(4\alpha Q({x}) - {\displaystyle \frac {1}{2 }} Q({x})^{3} - 2\, Q({x})\,{a} - {\displaystyle \frac {1}{2}}\,{\displaystyle \frac {{c}^{2}}{ Q({x})}} + {\displaystyle \frac {1}{2}}\, {\displaystyle \frac { {Q'({x})}^{2}}{ Q({x}) }} \nonumber\\ & &\quad + \tau'(x) + 2\,{x}^{2}\, Q({x}) - 2\,\tau({x})\, Q({x}) - 2\, Q({x}) + 2\, Q'(x) \,{x} \Biggr)\rho(x)\nonumber\\ & &\quad\quad - 2\, Q({x})\,{ \tau}(x) {x} - 2\, Q({x})^{2}\,{a} - {\displaystyle \frac {1 }{2}}\,{ \tau}(\,{x}\,)^{2} + {\displaystyle \frac {1}{2}}\,{ \gamma}^{2} - Q'({x}) \,{ \tau}(\,{x}\,)\nonumber \\ & &\quad\quad + Q({x})\, \tau'({x}) + 2\,{ \alpha }\,Q({x})^{2} - 2\,{ \tau}(\,{x}\,)\, Q({x})^{ 2} - {\displaystyle \frac {1}{2}}\,{c}^{2}=0.\label{p4minus} \end{eqnarray} That we obtain a quadratic equation in $\rho$ is a second difference between our results for \PII\ and \PIV . In general (beginning with a polynomial ODE that gives a Riccati equation for $\rho$) we could obtain at this stage a polynomial of higher degree in $\rho$. This quadratic equation must be compatible with the differential equation \begin{equation} \rho'(x)=-\rho(x)^2-(2x+2Q(x))\rho(x)+{\tau(x)}, \label{p4ric} \end{equation} for $\rho(x)$. Since equation (\ref{p4minus}) is quadratic in $\rho$, it is worth noting how we obtain a unique $\rho$ to check the compatibility condition. Suppose we differentiate Eqn (\ref{p4minus}) w.r.t. $x$ and use Eqn (\ref{p4ric}) to replace $\rho'$. Then we obtain a cubic equation in $\rho$. The third degree term can be eliminated by using Eqn (\ref{p4minus}) multiplied by $\rho$. This yields another quadratic equation for $\rho$. Eliminating the second-degree term by using Eqn (\ref{p4minus}) again, we get a linear equation for $\rho$, which we solve. The result is substituted into Eqn (\ref{p4ric}) to obtain the compatibility condition we investigate. This equation can be analysed to find $\tau$ and any conditions on the parameters. We assume here that $\tau$ is independent of both $Q$ and $Q'$; the final results are the same if we allow $\tau$ to depend on $Q$. The coefficient of $Q^{13}$ gives \begin{equation} \tau(x)={2\over 3}\,(\alpha - a). \label{tdag} \end{equation} Substitution of this into the compatibility condition yields \[ c^2=\gamma^2+{4\over 3}\,(\alpha^2-a^2-2\alpha + 2a). \] With this choice of $c^2$, the remaining terms in the compatibility condition factor to give \[ (\alpha -a)(2\alpha -6-3\gamma+4a)(2\alpha-6+3\gamma+4a)=0. \] Thus in addition to the identity $y(x)=Q(x)$ we obtain the two nontrivial BTs \begin{equation} \rho(x) = \,{(\pm\gamma-2+2\alpha)Q(x)\over Q'(x)+{Q(x)}^2+2xQ(x)\pm\gamma/2+1-\alpha} \label{rp41} \end{equation} \begin{eqnarray} a&=&-{1\over 4}\,(2\alpha - 6 \pm 3\gamma) \label{daga} \\ c^2&=&(\alpha \mp \gamma/2-1)^2 \label{dagc} \end{eqnarray} where the choice of sign of $\gamma$ (and $c$) arises because of the way we have written the second parameter in \PIV . This is the BT labelled by \lq\lq dagger\rq\rq\ ($y^{\dagger}$) in \cite{bch:95}. The alternative choice of sign in the dominant balance of terms in $\rho$ leads to the Riccati equation \[ \rho'(x)={\rho(x)}^2+\sigma(x)\rho(x), \] with \[ \sigma(x)=2Q(x)+2x+\tau(x)/\rho(x), \] where \[ \tau(x) = {2\over 3}\,(a-\alpha). \] This leads to the two BTs \[ \rho(x) = {(\pm\gamma-2-2\alpha)Q(x)\over \,Q'(x)-{Q(x)}^2-2xQ(x)\pm\gamma/2+1+\alpha} \] \begin{eqnarray*} a&=&-{1\over 4}\,(2\alpha + 6 \mp 3\gamma)\\ c^2&=&(\alpha \pm \gamma/2+1)^2 \end{eqnarray*} This is the BT labelled by \lq\lq double dagger\rq\rq ($y^{\ddagger}$) in \cite{bch:95}. It is interesting to note that all the known BTs \cite{bch:95} of \PIV\ can be expressed in terms of only two fundamental BTs. In \cite{bch:95}, these two BTs are labelled by \lq\lq hat\rq\rq ($\hat{y}$) and \lq\lq tilde\rq\rq ($\widetilde{y}$). In the next section, we show that our derivation of the dagger and double dagger BTs also gives rise to these two fundamental BTs. \section{Related BTs: the ODEs satisfied by {$\mathbf \rho(x)$}} The results of the previous Section were obtained by searching for auto-B\"acklund\ transformations. That is, we looked for transformations between two copies of the same equation distinguished by possibly different parameter values. In this Section, we consider the ODE satisfied by the function $\rho(x)$ in (\ref{trunc}), and deduce the B\"acklund\ transformation between this ODE and the Painlev\'e\ equation. \subsection{\PII } In constructing our first auto-BT for \PII (see the previous Section), we obtained a solution $\rho$ of the Riccati equation \[ \rho'(x) = \rho^2+2Q(x)\rho(x) \] given by \[ \rho(x) = {1-2\alpha \over 2{Q(x)}^2-2Q'(x)+x}, \] where $Q(x)$ satisfies \PII\ with parameter $a=-\alpha+1$. Now we eliminate $Q(x)$ from the above equations, to find a second order ODE satisfied by $\rho(x)$, together with B\"acklund\ transformations to \PII\ (in both $y(x)$ and $Q(x)$). Eliminating $Q(x)$ between the above two equations gives \[ \rho''(x) = \frac{3}{2}{\rho'(x)^2 \over \rho(x)} +\frac{1}{2} \rho(x)^3 +x\rho(x)-(1-2\alpha) \] which under $\rho(x)=1/s(x)$ becomes \[ s''(x) = \frac{1}{2}{s'(x)^2 \over s(x)} +(1-2\alpha)s(x)^2 -xs(x) -\frac{1}{2s(x)} \] which is \PtIV\ (i.e. the thirty-fourth equation in the classification results presented in Chapter 14 of \cite{ince}). Using this change of variables, and the expression for $Q$ in terms of $\rho$, the BT (\ref{trunc}) becomes \[ y(x) = {1-s'(x) \over 2s(x)}. \] Therefore, we recover the well-known mapping between \PtIV\ and \PII . We also have, of course, a B\"acklund\ transformation from \PtIV\ to \PII\ in $Q(x)$. If we had started with the second B\"acklund\ transformation for \PII , we again find (after a simple change of variables) the mapping to \PtIV . \subsection{\PIV } We now consider the ODE satisfied by $\rho(x)$ in our construction of B\"acklund\ transformations for \PIV . Recall from Equations (\ref{p4ric}), (\ref{tdag}) and (\ref{daga}), that $\rho(x)$ satisfies the Riccati equation \begin{equation} \rho'(x) = -\rho^2-2(Q(x)+x)\rho(x) +(\alpha-1\pm\gamma/2), \label{dagRic} \end{equation} for the dagger transformation. Elimination of $Q(x)$ between this and (\ref{rp41}) gives, after the substitution $\rho(x)=\pm C/s(x)$, the ODE \begin{equation} s''(x) = \frac{1}{2} {s'(x)^2 \over s(x)} +\frac{3}{2} s(x)^3 +4xs(x)^2 +2(x^2-A) s(x) -\frac{C^2}{2s(x)} \label{P4a} \end{equation} where \begin{eqnarray*} A & = & -\frac{1}{2} -\frac{1}{2}\alpha \pm\frac{3}{4}\gamma \\ \pm C & = & 1 -\alpha \mp\frac{1}{2}\gamma. \end{eqnarray*} This is another copy of the same equation, \PIV . The corresponding auto-B\"acklund\ transformation is obtained from (\ref{trunc}) after eliminating $Q(x)$ by using Equation (\ref{dagRic}). The result is \[ y(x) = {s'(x)-s(x)^2-2xs(x)+(1 -\alpha \mp\gamma/2) \over 2s(x)}. \] This is the BT labelled by \lq\lq tilde\rq\rq\ ($\widetilde{y}$) in \cite{bch:95}. On the other hand, Equation (\ref{rp41}) gives \[ s(x) = -{Q'(x)+Q(x)^2+2xQ(x)+(1+A\pm C/2) \over 2Q(x)} \] where by replacing $\alpha$, $\gamma$ in terms of $A$, $C$ above, we get the parameters $a$, $c$ of the version of \PIV\ satisfied by $Q(x)$ as \begin{eqnarray*} a & = & \frac{1}{2} -\frac{1}{2}A \pm\frac{3}{4}C \\ c^2 & = & \frac{1}{4}\left(2+2A\pm C\right)^2. \end{eqnarray*} Hence we get another BT for \PIV . This is the BT labelled by \lq\lq hat\rq\rq\ ($\hat{y}$) in \cite{bch:95}. That is, we have used our derivation of $y^{\dagger}$ to deduce both $\widetilde{y}$ and $\hat{y}$. The BTs $\widetilde{y}$ and $\hat{y}$ are fundamental BTs in the sense that all known BTs for \PIV\ can be expressed in terms of these two. (See \cite{bch:95}.) Our approach shows that $y^{\dagger}$ which maps $Q(x)$ to $y(x)$ can be considered, with appropriate choices of signs, as the composition $\widetilde{y}\circ\hat{y}$. Now consider our derivation of the double dagger BT in Section 2.2. Following the same procedure of elimination of $Q(x)$, we find (with the substitution $\rho=\mp C/s(x)$) the same \PIV\ i.e.\ Equation (\ref{P4a}), but with $A$ and $C$ given now by \begin{eqnarray*} A & = & \frac{1}{2} -\frac{1}{2}\alpha \mp\frac{3}{4}\gamma \\ \pm C & = & 1 +\alpha \mp\frac{1}{2}\gamma. \end{eqnarray*} The corresponding BT for \PIV\ is \[ y(x) = -{s'(x)+s(x)^2+2xs(x)+(1+\alpha\mp\gamma/2) \over 2s(x)}. \] This is $\hat{y}$. Elimination of $y(x)$ instead gives \[ s(x) = {Q'(x)-Q(x)^2-2xQ(x)+(1-A\pm C/2) \over 2Q(x)} \] where \begin{eqnarray*} a & = & -\frac{1}{2} -\frac{1}{2}A \mp\frac{3}{4}C \\ c^2 & = & \frac{1}{4}\left(2-2A\pm C\right)^2, \end{eqnarray*} which is $\widetilde{y}$. Thus our derivation of $y^{\ddagger}$ also allows us to obtain the fundamental BTs $\widetilde{y}$ and $\hat{y}$. Our approach then shows that, for suitable choices of signs, $y^{\ddagger}$ can be expressed in terms of $\widetilde{y}$ and $\hat{y}$ as the composition $\hat{y}\circ\widetilde{y}$ \cite{bch:95}. \section{The double-singularity approach} In Sections 2 and 3, we assumed that $\rho$ inherited half the poles of the Painlev\'e\ transcendent $y(x)$. Now we consider the possibility that both families of movable poles of $y(x)$ are inherited by specified functions called respectively $\rho_1$ and $\rho_2$ in the transformation. That is, we rewrite $y$ as \begin{equation} y(x) = \rho_1(x) - \rho_2(x) + Y(x), \label{double} \end{equation} where we assume that $Y(x)$ satisfies \PII , or respectively \PIV , with different parameters ($a$, or respectively $a$ and $c$). As before, dominant terms of \PII\ or \PIV\ lead to Riccati equations for $\rho_i$ to leading-order near a movable pole. The dominant terms in each equation are as before. However, the lower-order terms cannot be uniquely determined by dominant balances. For simplicity, we choose the Riccati equation in $\rho_1$, to be linear in $\rho_2$ and {\it vice versa}\/. Consider the Riccati equations for $\rho_i$ to have the form \begin{eqnarray} {\rho_1}'(x)&=& {\rho_1(x)}^2+A_1(x)\rho_1(x)\rho_2(x)+B_1(x)\rho_1(x) \nonumber\\ & &\quad +C_1(x)\rho_2(x)+\tau_1(x)\label{r1}\\ {\rho_2}'(x)&=& {\rho_2(x)}^2+A_2(x)\rho_1(x)\rho_2(x)+B_2(x)\rho_1(x) \nonumber\\ & &\quad +C_2(x)\rho_2(x)+\tau_2(x)\label{r2} \end{eqnarray} In this case, where we take the same sign against $\rho_i^2$ in each Riccati equation, this approach is analagous to the double singular manifold method. However, we show at the end of each subsection below that we can also obtain nontrivial results by taking opposite signs in the above Riccati equations. \subsection{\PII } We transform \PII\ by using the relation (\ref{double}) and equations (\ref{r1}), (\ref{r2}). Then we get \begin{eqnarray} & &- { \alpha} +a-{C}_{2}(x){ \tau}_{2}(x)+{\tau_1}'(x) +{B}_{1}(x){ \tau}_{1}(x)-{B}_{2}(x){ \tau}_{1}(x)\nonumber\\ & &\quad + {C}_{1}(x){ \tau}_{2}(x)-{\tau_2}'(x)\nonumber\\ & &\quad + \Bigl(3B_1(x)-B_2(x)-6Y(x)-A_2(x)B_2(x)+A_1(x)B_2(x)\Bigr) {\rho_1(x)}^2\nonumber\\ & &\quad + \Bigl({B_1}'(x)-B_2(x)B_1(x)+C_1(x)B_2(x)-{B_2}'(x)\nonumber\\ & &\quad\quad -6{Y(x)}^2 + A_1(x){ \tau}_{2}(x)+{B_1(x)}^2-C_2(x)B_2(x) +2\tau_1(x)-x\nonumber\\ & &\quad\quad -A_2(x)\tau_2(x)\Bigr)\rho_1(x)+\Bigl(C_1(x)C_2(x)+ A_1(x){ \tau}_{1}(x)+6{Y(x)}^2 \nonumber\\ & &\quad\quad +B_1(x)C_1(x) -{C_2(x)}^2+x +{C_1}'(x)-{C_2}'(x)-2\tau_2(x)\nonumber\\ & &\quad\quad -C_1(x)B_2(x)-A_2(x)\tau_1(x)\Bigr)\rho_2(x)\nonumber\\ & &\quad +\Bigl(C_1(x)-6Y(x)-3C_2(x)-C_1(x)A_2(x)+A_1(x)C_1(x)\Bigr) {\rho_2(x)}^2 \nonumber\\ & &\quad +\Bigl(-A_2(x)+A_1(x)A_2(x)+3A_1(x)+6-{A_2(x)}^2\Bigr) {\rho_1(x)}^2\rho_2(x)\nonumber\\ & &\quad +\Bigl(A_1(x)-A_1(x)A_2(x)-3A_2(x)-6+{A_1(x)}^2\Bigr) \rho_1(x){\rho_2(x)}^2\nonumber\\ & &\quad +\Bigl(2C_1(x)-{A_2}'(x)-2B_2(x)+A_1(x)C_2(x)+C_1(x)A_2(x)\nonumber\\ & &\quad \quad +12 Y(x)-2C_2(x)A_2(x)+2B_1(x)A_1(x)-A_1(x)B_2(x)\nonumber\\ & &\quad \quad +{A_1}'(x)-A_2(x)B_1(x)\Bigr)\rho_1(x)\rho_2(x)=0.\label{p2_double} \end{eqnarray} The dominant balances of this equation give \begin{eqnarray} -A_2(x)+A_1(x)A_2(x)+3A_1(x)+6-{A_2(x)}^2&=&0\label{112}\\ 3B_1(x)-B_2(x)-6Y(x)-A_2(x)B_2(x)+A_1(x)B_2(x)&=&0\label{11}\\ A_1(x)-A_1(x)A_2(x)-3A_2(x)-6+{A_1(x)}^2&=&0\label{122}\\ C_1(x)-6Y(x)-3C_2(x)-C_1(x)A_2(x)+A_1(x)C_1(x)&=&0\label{22} \end{eqnarray} The sum of Eqns (\ref{112}) and (\ref{122}) factors to give \[ (A_2-A_1)(A_1+A_2+4)=0. \] Substitution of the first solution $A_2=A_1$ shows that \begin{equation} A_2 = A_1 = -3 \label{p2_case1} \end{equation} On the other hand, the second solution $A_1=-A_2-4$ gives \[ A_1=-1, A_2=-3\ \textrm{or}\ A_1 = -3, A_2 = -1. \] By relabelling if necessary, we take this case to be \begin{equation} A_1= -1, A_2= -3. \label{p2_case2} \end{equation} \subsubsection{Case 1: $A_2 = A_1 = -3$} Consider the first case (\ref{p2_case1}). Equations (\ref{11}) and (\ref{22}) give \[ 3B_1=B_2+6Y,\quad C_1=3C_2+6Y. \] We use these to replace $B_1$, $C_1$ in the transformed equation (\ref{p2_double}). Now the dominant terms give \begin{eqnarray} 3\tau_1&=&{B_2}'(x)+{3x\over 2}+{{B_2(x)}^2\over 3}-8B_2(x)Y(x)+3Y(x)^2\nonumber\\ &&\quad -3C_2(x)B_2(x)-3Y'(x)\label{c11}\\ \tau_2&=&{C_2}'(x)+{x\over 2}+{C_2(x)}^2-2B_2(x)Y(x)+9Y(x)^2\nonumber\\ &&\quad +6C_2(x)Y(x)+3Y'(x)\label{c12}-C_2(x)B_2(x) \end{eqnarray} The transformed equation then becomes \begin{eqnarray*} & & -\alpha-3a+\Bigl(C_2(x)-{\displaystyle\frac{1}{3}}B_2(x)\Bigr)x -2C_2(x)^2 B_2(x) +2C_2(x)^3 -{\displaystyle\frac{2}{27}} B_2(x)^3 \nonumber \\ & & +{\displaystyle\frac{2}{3}} C_2(x) B_2(x)^2 +2 B_2(x)^2 Y(x) +18 C_2(x)^2 Y(x) -12 C_2(x) B_2(x) Y(x) \nonumber \\ & & +54 C_2(x) Y(x)^2 -18 B_2(x) Y(x)^2 +{\displaystyle\frac{1}{3}} B_2''(x) - C_2''(x) = 0, \end{eqnarray*} which, under the change of variables \[ B_2(x)=3C_2(x)+3V(x)+9Y(x) \] becomes \[ V''(x)-2V(x)^3-xV(x)-\alpha=0, \] which is just \PII\ with parameter $\alpha$ (the same parameter as that in the version of \PII\ satisfied by $y(x)$). These results for the coefficients $B_1$, $C_1$, $B_2$, motivate the change of variables \begin{eqnarray*} \rho_1(x)&=&C_2(x)+2Y(x)+\sigma_1(x) \\ \rho_2(x)&=&C_2(x)+3Y(x)+V(x)+\sigma_2(x) \end{eqnarray*} which simplify the Riccati equations for $\rho_1$ and $\rho_2$. These become \begin{eqnarray} {\sigma_1}'(x)&=&{\sigma_1(x)}^2-3\sigma_1(x)\sigma_2(x)\nonumber\\ & &\quad -2\sigma_1(x)V(x)+V'(x)+V(x)^2+x/2\label{sig1}\\ {\sigma_2}'(x)&=&{\sigma_2(x)}^2-3\sigma_1(x)\sigma_2(x)\nonumber\\ & &\quad +2\sigma_2(x)V(x)-V'(x)+V(x)^2+x/2\label{sig2} \end{eqnarray} We can eliminate $\sigma_2(x)$ from this system by solving the first equation for $\sigma_2(x)$ and substituting the result into the second. The result is, for $W(x)=2\sigma_1(x)$, \begin{eqnarray} W''(x)&=&{2\over 3}\,{{W'(x)}^2\over W(x)}-{\displaystyle\frac{1}{3}} \left(2W(x)-2V(x)+{2V'(x)+x+2V(x)^2\over W(x)}\right)W'(x)\nonumber\\ & &\quad +{\displaystyle\frac{2}{3}}{W(x)}^3 -{\displaystyle\frac{10}{3}}V(x){W(x)}^2 +{\displaystyle\frac{1}{3}}\left(6V(x)^2+10V'(x)-x\right)W(x)\nonumber\\ & &\quad +{\displaystyle\frac{8}{3}}V(x)^3 +{\displaystyle\frac{4}{3}}xV(x)+1 +{\displaystyle\frac{8}{3}}V(x)V'(x)+2\alpha -{\displaystyle\frac{1}{3}}\left(x^2+4V(x)^4 \right.\nonumber\\ & &\quad \left. +4xV'(x)+8V(x)^2V'(x) +4xV(x)^2+4{\bigl(V'(x)\bigr)}^2\right){1\over W(x)}\label{w35} \end{eqnarray} This is \PtV\ whose general form (where the functions $r(x)$ and $q(x)$ are as given in \cite{ince}) is \begin{eqnarray*} {d^2W\over dx^2}&=&{2\over 3W}\left({dW\over dx}\right)^2 -\left({2W\over 3}-{2q\over 3}-{r\over W}\right){dW\over dx}\\ &&\quad +{2W^3\over 3}-{10\over 3}qW^2+\left(4q'+r+8q^2/3\right)W\\ &&\quad +2qr-3r'-{3r^2\over W}. \end{eqnarray*} Equation (\ref{w35}) is this equation with choices of coefficients given by \[ q(x)=V(x),\quad r(x)=-{2\over 3}\Bigl(V'(x)+V(x)^2\Bigr)-\frac{1}{3}x . \] (We could have eliminated $\sigma_1$ in the above instead of $\sigma_2$. The result is the same except for a sign change: $q=-V$.) Writing the B\"acklund\ transformation in terms of $W(x)=2\sigma_1(x)$ gives \begin{eqnarray*} y(x)&=& \sigma_1(x)-\sigma_2(x)-V(x)\\ &=& {\displaystyle\frac{1}{3}}W(x) -{\displaystyle\frac{1}{3}}{x \over W(x)} +{\displaystyle\frac{1}{3}}{W'(x) \over W(x)} -{\displaystyle\frac{2}{3}}{V(x)^2 \over W(x)} -{\displaystyle\frac{1}{3}} V(x) -{\displaystyle\frac{2}{3}}{V'(x) \over W(x)} \end{eqnarray*} with inverse \[ W(x) = {y'(x)-V'(x)+y(x)^2-V(x)^2 \over y(x)-V(x)} . \] Hence we have reproduced the relation between two solutions of \PII\ and a solution of \PtV\ that has been known since classical studies of these equations \cite{ince}. \subsubsection{Case 2: $A_1= -1, A_2= -3.$} Now we consider the second case of possible values of $A_1$, $A_2$ given by Eqn (\ref{p2_case2}). Here we find from (\ref{11}) and (\ref{22}) that \begin{eqnarray*} C_1(x) & = & C_2(x)+2Y(x) \\ B_1(x) & = & -{\displaystyle\frac{1}{3}}B_2(x)+2Y(x) . \end{eqnarray*} Substituting these values of $C_1(x)$ and $B_1(x)$ into (\ref{p2_double}) gives (from the coefficient of $\rho_1(x)\rho_2(x)$) \[ B_2(x) = 3C_2(x)+9Y(x). \] The transformed equation then gives \begin{eqnarray*} \tau_1(x) & = & C_2'(x)-C_2(x)Y(x)-2Y(x)^2+2Y'(x) \\ \tau_2(x) & = & C_2'(x)-2C_2(x)^2-9C_2(x)Y(x)-9Y(x)^2+3Y'(x)+\frac{1}{2}x \end{eqnarray*} and substitution of these values of $\tau_1(x)$ and $\tau_2(x)$ gives \[ \alpha = -\,1/2. \] Our subsequent results apply only to this special case of \PII . Making the change of variables \begin{eqnarray*} \rho_1(x)&=&C_2(x)+2Y(x)+\sigma_1(x)\\ \rho_2(x)&=&C_2(x)+3Y(x)+\sigma_2(x) \end{eqnarray*} we find that the BT becomes \[ y(x)=\sigma_1(x) - \sigma_2(x). \] Note that $Y$ and $a$ have been eliminated here. The equations satisfied by $\sigma_1(x)$ and $\sigma_2(x)$ now yield two (apparently) different second-order equations. Elimination of $\sigma_2(x)$ gives \[ \sigma_1''(x)-2\sigma_1(x)^3+\frac{1}{2}x\sigma_1(x)=0 \] which is equivalent to \PII\ with zero value of the parameter. The BT in this case is \[ y(x)={{\sigma_1}'(x)\over \sigma_1(x)}, \] where $y(x)$ satisfies \PII\ with $\alpha = -\,1/2$. This BT, from \PII\ with $\alpha = 0$ to \PII\ with $\alpha = -\,1/2$, does not seem to be widely known, although it can in fact be found in \cite{gambier:acta}. On the other hand, elimination of $\sigma_1(x)$ gives (with $\sigma_2(x)=-W(x)$) \begin{eqnarray} W''(x) & = & \frac{2}{3}{(W'(x))^2 \over W(x)} -\frac{1}{3} \left(2W(x)-{x \over 2W(x)} \right)W'(x) +\frac{2}{3} W(x)^3 \nonumber\\ & & +\frac{1}{6}xW(x)-\frac{1}{12} {x^2 \over W(x)} -\frac{1}{2} \label{ff1} \end{eqnarray} which can be rescaled onto \PtV\ with $q=0$, $r=-x/3$. The BT in this case is \begin{equation} y(x)= {2\over 3}W(x) - {W'(x)\over 3W(x)} -{x\over 6W(x)}. \label{ff2} \end{equation} We now briefly consider the results obtained from our double-singularity approach if we assume our Riccati system to be \begin{eqnarray} {\rho_1}'(x)&=& {\rho_1(x)}^2+A_1(x)\rho_1(x)\rho_2(x)+B_1(x)\rho_1(x) \nonumber\label{rr1}\\ & &\quad +C_1(x)\rho_2(x)+\tau_1(x)\\ {\rho_2}'(x)&=& -{\rho_2(x)}^2+A_2(x)\rho_1(x)\rho_2(x)+B_2(x)\rho_1(x) \nonumber\label{rr2}\\ & &\quad +C_2(x)\rho_2(x)+\tau_2(x) \end{eqnarray} With this choice of Riccati system we obtain two inequivalent choices of coefficients $A_1(x)$, $A_1(x)$. These are: \[ A_1(x)=3,\quad A_2(x)=-3, \qquad {\rm or} \qquad A_2(x)=A_1(x)+2. \] With the first of these choices our final results are that under the change of variables \begin{eqnarray*} \rho_1(x) & = & {\displaystyle\frac{1}{3}}B_2(x)-Y(x)+\sigma_1(x) \\ \rho_2(x) & = & {\displaystyle\frac{1}{3}}B_2(x)+\sigma_2(x) \end{eqnarray*} our B\"acklund\ transformation becomes \[ y(x)=\sigma_1(x)-\sigma_2(x), \] where $\sigma_1(x)$ and $\sigma_2(x)$ satisfy the Riccati system \begin{eqnarray*} \sigma_1'(x) & = & \sigma_1(x)^2+3\sigma_1(x)\sigma_2(x) -{\displaystyle\frac{1}{4}}x \\ \sigma_2'(x) & = & -\sigma_2(x)^2-3\sigma_1(x)\sigma_2(x) +{\displaystyle\frac{1}{4}}x \end{eqnarray*} and we have in addition the compatibility condition (resulting from the transformed equation) \[ \alpha=-\frac{1}{2} \] Note that again $Y$ and $a$ play no role in our final results. Eliminating $\sigma_2(x)$ from the above system and then substituting $\sigma_1(x)=W(x)/2$ yields equation (\ref{ff1}), together with the same B\"ack\-lund transformation (\ref{ff2}). On the other hand, eliminating $\sigma_1(x)$ from the above Riccati system yields (\ref{ff1}) in $W(x)=-2\sigma_2(x)$. With the second choice of coefficients $A_i(x)$ we find only that $y(x)=u(x)$ satisfies \PII\ for $\alpha=1/2$, where $u(x)=\rho_1(x)-\rho_2(x)+Y(x)$ is a solution of \[ u'(x)=u(x)^2+\frac{1}{2}x \] i.e.\ we find only the special integral of \PII\ which gives rise to Airy function solutions \cite{gambier:acta}. \subsection{\PIV } Dominant balances of the transformed \PIV\ equation yield only one consistent choice of $A_1$, $A_2$, namely \[ A_1(x)=A_2(x)=-2. \] The resulting equation shows that \[ C_2(x)=-2x-2Y(x),\quad B_1(x)=2x+2Y(x),\quad C_1(x)=-4x-2Y(x)+B_2(x), \] and that \begin{eqnarray*} \tau_1(x) & = & {\displaystyle\frac{1}{2}} B_2'(x) -{\displaystyle\frac{1}{4}}B_2(x)^2 +xB_2(x)-Y'(x)+Y(x)^2+2xY(x) \\ & & -\alpha-1 \pm {\displaystyle\frac{1}{2}} \gamma \\ \tau_2(x) & = & {\displaystyle\frac{1}{2}} B_2'(x) -{\displaystyle\frac{1}{4}}B_2(x)^2 +xB_2(x)+Y(x)B_2(x)-\alpha-1 \mp {\displaystyle\frac{1}{2}} \gamma \end{eqnarray*} Defining $\rho_i$ now by taking \begin{eqnarray*} \rho_1(x)&=&{\displaystyle\frac{1}{2}}B_2(x)-Y(x)-2x+\sigma_1(x)\\ \rho_2(x)&=&{\displaystyle\frac{1}{2}}B_2(x)+\sigma_2(x), \end{eqnarray*} the BT becomes \[ y(x)=\sigma_1(x)-\sigma_2(x)-2x. \] Once again the intermediate variable $Y(x)$ (and the parameters $a$ and $c$) has been eliminated and we are now transforming to $y(x)$ from $\sigma_1$ or $\sigma_2$. Eliminating $\sigma_2(x)$ leads to a different version of \PIV\ \[ {\sigma_1}''(x)={{{\sigma_1}'(x)}^2\over 2\sigma_1(x)} +{3\over 2}{\sigma_1(x)}^3 - 4x{\sigma_1(x)}^2 +2(x^2-A){\sigma_1(x)}-{C^2\over 2\sigma_1(x)} \] where \begin{eqnarray*} A&=&-{1\over 2}\alpha\mp{3\over 4}\gamma-{1\over 2}\\ C^2&=&{1\over 4}(\pm\gamma-2\alpha+2)^2 \end{eqnarray*} (This is the conventional form of \PIV\ if we take $V=-\sigma_1$.) In this case, the BT becomes \begin{equation} y(x) = {{\sigma_1}'(x) + \sigma_1(x)^2 -2x\sigma_1(x) -(1-\alpha\pm\gamma/2) \over 2\sigma_1(x)}. \end{equation} Elimination of $\sigma_1(x)$ leads to a conventional version of \PIV\ governing $\sigma_2(x)$ but with new $A$ and $C$ given by \begin{eqnarray*} A&=&-{1\over 2}\alpha\pm{3\over 4}\gamma+{1\over 2}\\ C^2&=&{1\over 4}(\pm\gamma+2\alpha+2)^2 \end{eqnarray*} In this case, the BT is \begin{equation} y(x) = -{{\sigma_2}'(x) + \sigma_2(x)^2 +2x\sigma_2(x) +(1+\alpha\pm\gamma/2) \over 2\sigma_2(x)}. \end{equation} Thus once again we recover the BTs $\widetilde{y}$ and $\hat{y}$ obtained in Section 3. Considering also the BT from $\sigma_2$ to $\sigma_1$ (respectively $\sigma_1$ to $\sigma_2$) then gives, for suitable choices of signs, $\hat{y}= \widetilde{y}\circ\widetilde{y}$ (respectively $\widetilde{y}=\hat{y}\circ \hat{y}$) \cite{bch:95}. For \PIV , the Riccati system (\ref{rr1}), (\ref{rr2}) gives only restricted results. We get that $y(x)=u(x)$ is a solution of \PIV\ with parameters subject to the constraint $\gamma^2=4(\alpha+1)^2$, where $u(x)=\rho_1(x)-\rho_2(x)+Y(x)$ satisfies \[ u'(x)=u(x)^2+2xu(x)-2(\alpha+1), \] i.e.\ we find only a special integral of \PIV . \section{Conclusions} We have introduced a new, general method of constructing B\"acklund\ transformations for ordinary differential equations. This is based on mappings preserving natural subsets of movable poles. For the examples considered here this method has allowed us to construct all known fundamental B\"acklund\ transformations, including a less well known B\"acklund\ transformation for \PII . Our approach has also allowed us to find B\"acklund\ transformations onto other ODEs in the Painlev\'e classification, as well as to deduce relationships between the B\"acklund\ transformations constructed. Its application to other Painlev\'e equations is discussed in \cite{gjp2}. \section*{Acknowledgements} Andrew Pickering and Pilar R.\ Gordoa are grateful to Nalini Joshi for her invitations to visit the University of Adelaide. The research in this paper was supported by the Australian Research Council. \newpage
\section{Introduction} Einstein's gravitation theory can be thought of as defined by two postulates. One postulate states that the action functional describing the propagation and self-interaction of the gravitational field is \begin{equation} S_{\rm gravitation} = {c^4 \over 16\pi \ G} \int {d^4 x \over c} \ \sqrt{g} \ R(g). \label{eq:01} \end{equation} A second postulate states that the action functional describing the coupling of all the fields describing matter and its electro-weak and strong interactions (leptons and quarks, gauge and Higgs bosons) is a (minimal) deformation of the special relativistic action functional used by particle physicists (the so called ``Standard Model''), obtained by replacing everywhere the flat Minkowski metric $\eta_{\mu \nu} = {\rm diag} (-1,+1,+1,+1)$ by $g_{\mu \nu} (x^{\lambda})$ and the partial derivatives $\partial_{\mu} \equiv \partial / \partial x^{\mu}$ by $g$-covariant derivatives $\nabla_{\mu}$. Schematically, one has \begin{equation} S_{\rm matter} = \int { d^4 x \over c} \ \sqrt{g} \ {\cal L}_{\rm matter} \ [\psi ,A_{\mu},H;g_{\mu \nu}]. \label{eq:02.a} \end{equation} Einstein's theory of gravitation is then defined by extremizing the total action functional, $S_{\rm tot} \ [g,\psi ,A,H] = S_{\rm gravitation} \ [g] + S_{\rm matter} \ [\psi ,A,H,g]$. Although, seen from a wider perspective, the two postulates (\ref{eq:01}) and (\ref{eq:02.a}) follow from the unique requirement that the gravitational interaction be mediated only by massless spin-2 excitations \cite{1}, the decomposition in two postulates is convenient for discussing the theoretical significance of various tests of General Relativity. Let us discuss in turn the experimental tests of the coupling of matter to gravity (postulate (\ref{eq:02.a})), and the experimental tests of the dynamics of the gravitational field (postulate (\ref{eq:01})). For more details and references we refer the reader to \cite{2} or \cite{3}. \section{Experimental tests of the coupling between matter and gravity} The fact that the matter Lagrangian depends only on a symmetric tensor $g_{\mu \nu} (x)$ and its first derivatives (i.e. the postulate of a universal ``metric coupling'' between matter and gravity) is a strong assumption (often referred to as the ``equivalence principle'') which has many observable consequences for the behaviour of localized test systems embedded in given, external gravitational fields. In particular, it predicts the constancy of the ``constants'' (the outcome of local non-gravitational expe\-riments, referred to local standards, depends only on the values of the coupling constants and mass scales entering the Standard Model) and the universality of free fall (two test bodies dropped at the same location and with the same velocity in an external gravitational field fall in the same way, independently of their masses and compositions). Many sorts of data (from spectral lines in distant galaxies to a natural fission reactor phenomenon which took place at Oklo, Gabon, two billion years ago) have been used to set limits on a possible time variation of the basic coupling constants of the Standard Model. The best results concern the electromagnetic coupling, i.e. the fine-structure constant $\alpha_{\rm em}$. A recent reanalysis of the Oklo phenomenon gives a conservative upper bound \cite{5} \begin{equation} -6.7 \times 10^{-17} \, {\rm yr}^{-1} < {{\dot{\alpha}}_{\rm em} \over \alpha_{\rm em}} < 5.0 \times 10^{-17} \ {\rm yr}^{-1} , \label{eq:05} \end{equation} which is much smaller than the cosmological time scale $\sim 10^{-10} \ {\rm yr}^{-1}$. It would be interesting to confirm and/or improve the limit (\ref{eq:05}) by direct laboratory measurements comparing clocks based on atomic transitions having different dependences on $\alpha_{\rm em}$. [Current atomic clock tests of the constancy of $\alpha_{\rm em}$ give the limit $\vert {\dot{\alpha}}_{\rm em} / \alpha_{\rm em} \vert < 3.7 \times 10^{-14} \, {\rm yr}^{-1}$ \cite{PTM95}.] The universality of free fall has been verified at the $10^{-12}$ level both for laboratory bodies \cite{7}, e.g. (from the last reference in \cite{7}) \begin{equation} \left( \frac{\Delta a}{a} \right)_{\rm Be \, Cu} = (-1.9 \pm 2.5) \times 10^{-12} \, , \label{eq:00.a} \end{equation} and for the gravitational accelerations of the Moon and the Earth toward the Sun \cite{8}, \begin{equation} \left(\frac{\Delta a}{a}\right)_{\rm Moon \, Earth} = (-3.2 \pm 4.6) \times 10^{-13} \, . \label{eq:00.b} \end{equation} In conclusion, the main observable consequences of the Einsteinian postulate (\ref{eq:02.a}) concerning the coupling between matter and gravity (``equivalence principle'') have been verified with high precision by all experiments to date (see Refs. \cite{2}, \cite{3} for discussions of other tests of the equivalence principle). The traditional paradigm (first put forward by Fierz \cite{10}) is that the extremely high precision of free fall experiments ($10^{-12}$ level) strongly suggests that the coupling between matter and gravity is exactly of the ``metric'' form (\ref{eq:02.a}), but leaves open possibilities more general than eq. (\ref{eq:01}) for the spin-content and dynamics of the fields mediating the gravitational interaction. We shall provisionally adopt this paradigm to discuss the tests of the other Einsteinian postulate, eq. (\ref{eq:01}). However, we shall emphasize at the end that recent theoretical findings suggest a new paradigm. \section{Tests of the dynamics of the gravitational field in the weak field regime} Let us now consider the experimental tests of the dynamics of the gravitational field, defined in General Relativity by the action functional (\ref{eq:01}). Following first the traditional paradigm, it is convenient to enlarge our framework by embedding General Relativity within the class of the most natural relativistic theories of gravitation which satisfy exactly the matter-coupling tests discussed above while differing in the description of the degrees of freedom of the gravitational field. This class of theories are the metrically-coupled tensor-scalar theories, first introduced by Fierz \cite{10} in a work where he noticed that the class of non-metrically-coupled tensor-scalar theories previously introduced by Jordan \cite{11} would generically entail unacceptably large violations of the equivalence principle. The metrically-coupled (or equivalence-principle respecting) tensor-scalar theories are defined by keeping the postulate (\ref{eq:02.a}), but replacing the postulate (\ref{eq:01}) by demanding that the ``physical'' metric $g_{\mu \nu}$ (coupled to ordinary matter) be a composite object of the form \begin{equation} g_{\mu \nu} = A^2 (\varphi) \ g_{\mu \nu}^* , \label{eq:06} \end{equation} where the dynamics of the ``Einstein'' metric $g_{\mu \nu}^*$ is defined by the action functional (\ref{eq:01}) (written with the replacement $g_{\mu \nu} \rightarrow g_{\mu \nu}^*$) and where $\varphi$ is a massless scalar field. [More generally, one can consider several massless scalar fields, with an action functional of the form of a general nonlinear $\sigma$ model \cite{12}]. In other words, the action functional describing the dynamics of the spin 2 and spin 0 degrees of freedom contained in this generalized theory of gravitation reads \begin{eqnarray} S_{\rm gravitational} \, [g_{\mu \nu}^* ,\varphi ] \!\!\!\!\!\!&=\!\!\!\!\!& {c^4 \over 16\pi \, G_*} \int {d^4 x \over c} \, \sqrt{g_*} \nonumber \\ &\times\!\!\!\!\!&[R(g_*) - 2g_*^{\mu \nu} \, \partial_{\mu} \, \varphi \, \partial_{\nu} \, \varphi ] . \label{eq:07} \end{eqnarray} Here, $G_*$ denotes some bare gravitational coupling constant. This class of theories contains an arbitrary function, the ``coupling function'' $A(\varphi)$. When $A(\varphi) = {\rm const.}$, the scalar field is not coupled to matter and one falls back (with suitable boundary conditions) on Einstein's theory. The simple, one-parameter subclass $A(\varphi) = \exp (\alpha_0 \ \varphi)$ with $\alpha_0 \in {\bb R}$ is the Jordan-Fierz-Brans-Dicke theory \cite{10}, \cite{J59}, \cite{BD}. In the general case, one can define the (field-dependent) coupling strength of $\varphi$ to matter by \begin{equation} \alpha (\varphi) \equiv {\partial \ln A(\varphi) \over \partial \varphi} . \label{eq:08} \end{equation} It is possible to work out in detail the observable consequences of tensor-scalar theories and to contrast them with the general relativistic case (see, e.g., ref. \cite{12}). Let us now consider the experimental tests of the dynamics of the gravitational field that can be performed in the solar system. Because the planets move with slow velocities $(v/c \sim 10^{-4})$ in a very weak gravitational potential $(U/c^2 \sim (v/c)^2 \sim 10^{-8})$, solar system tests allow us only to probe the quasi-static, weak-field regime of relativistic gravity (technically described by the so-called ``post-Newtonian'' expansion). In the limit where one keeps only the first relativistic corrections to Newton's gravity (first post-Newtonian approximation), all solar-system gravitational experiments, interpreted within tensor-scalar theories, differ from Einstein's predictions only through the appearance of two ``post-Einstein'' parameters $\overline{\gamma}$ and $\overline{\beta}$ (related to the usually considered Eddington parameters $\gamma$ and $\beta$ through $\overline{\gamma} \equiv \gamma -1$, $\overline{\beta} \equiv \beta -1$). The parameters $\overline{\gamma}$ and $\overline{\beta}$ vanish in General Relativity, and are given in tensor-scalar theories by \begin{equation} \overline{\gamma} = -2 \ {\alpha_0^2 \over 1+\alpha_0^2} , \label{eq:09.a} \end{equation} \begin{equation} \overline{\beta} = +{1 \over 2} \ {\beta_0 \ \alpha_0^2 \over (1+\alpha_0^2)^2} , \label{ref:09.b} \end{equation} where $\alpha_0 \equiv \alpha (\varphi_0)$, $\beta_0 \equiv \partial \alpha (\varphi_0) / \partial \varphi_0$; $\varphi_0$ denoting the cosmologically-determined value of the scalar field far away from the solar system. Essentially, the parameter $\overline{\gamma}$ depends only on the linearized structure of the gravitational theory (and is a direct measure of its field content, i.e. whether it is pure spin 2 or contains an admixture of spin 0), while the parameter $\overline{\beta}$ parametrizes some of the quadratic nonlinearities in the field equations (cubic vertex of the gravitational field). All currently performed gravitational experiments in the solar system, including perihelion advances of planetary orbits, the bending and delay of electromagnetic signals passing near the Sun, and very accurate range data to the Moon obtained by laser echoes, are compatible with the general relativistic predictions $\overline{\gamma} = 0 =\overline{\beta}$ and give upper bounds on both $\left\vert \overline{\gamma} \right\vert$ and $\left\vert \overline{\beta} \right\vert$(i.e. on possible fractional deviations from General Relativity). The best current limits come from: (i) VLBI measurements of the deflection of radio waves by the Sun, giving \cite{eubanks}: $-3.8 \times 10^{-4} < \overline{\gamma} < 2.6 \times 10^{-4}$, and (ii) Lunar Laser Ranging measurements of a possible polarization of the orbit of the Moon toward the Sun (``Nordtvedt effect'' \cite{N68}) giving \cite{8}: $4\overline{\beta} - \overline{\gamma} = -0.0007 \pm 0.0010$. The corresponding bounds on the scalar coupling parameters $\alpha_0$ and $\beta_0$ are: $\alpha_0^2 < 1.9 \times 10^{-4}$, $-8.5 \times 10^{-4} < (1+\beta_0) \alpha_0^2 < 1.5 \times 10^{-4}$. Note that if one were working in the more general (and more plausible; see below) framework of theories where the scalar couplings violate the equivalence principle one would get much stronger constraints on the basic coupling parameter $\alpha_0$ of order $\alpha_0^2 \lsim 10^{-7}$ \cite{dv96a}. The parametrization of the weak-field deviations between generic tensor-scalar theories and Einstein's theory has been extended to the second post-Newto\-nian order \cite{14}. Only two post-post-Einstein parameters, $\varepsilon$ and $\zeta$, representing a deeper layer of structure of the gravitational interaction, show up. These parameters have been shown to be already significantly constrained by binary-pulsar data: $\vert \varepsilon \vert < 7 \times 10^{-2}$, $\vert \zeta \vert < 6 \times 10^{-3}$. \section{Tests of the dynamics of the gravitational field in the strong field regime} In spite of the diversity, number and often high precision of solar system tests, they have an important qualitative weakness : they probe neither the radiation pro\-perties nor the strong-field aspects of relativistic gravity. Fortunately, the discovery \cite{15} and continuous observational study of pulsars in gravitationally bound binary orbits has opened up an entirely new testing ground for relativistic gravity, giving us an experimental handle on the regime of strong and/or radiative gravitational fields. The fact that binary pulsar data allow one to probe the propagation properties of the gravitational field is well known. This comes directly from the fact that the finite velocity of propagation of the gravitational interaction between the pulsar and its companion generates damping-like terms in the equations of motion, i.e. terms which are directed against the velocities. [This can be understood heuristically by considering that the finite velocity of propagation must cause the gravitational force on the pulsar to make an angle with the instantaneous position of the companion \cite{16}, and was verified by a careful derivation of the general relativistic equations of motion of binary systems of compact objects \cite{17}]. These damping forces cause the binary orbit to shrink and its orbital period $P_b$ to decrease. The measurement, in some binary pulsar systems, of the secular orbital period decay $\dot{P}_b \equiv dP_b / dt$ \cite{18} thereby gives us a direct experimental probe of the damping terms present in the equations of motion. The fact that binary pulsar data allow one to probe strong-field aspects of re\-lativistic gravity is less well known. The a priori reason for saying that they should is that the surface gravitational potential of a neutron star $Gm / c^2 R \simeq 0.2$ is a mere factor 2.5 below the black hole limit (and a factor $\sim 10^8$ above the surface potential of the Earth). Due to the peculiar ``effacement'' properties of strong-field effects taking place in General Relativity \cite{17}, the fact that pulsar data probe the strong-gravitational-field regime can only be seen when contrasting Einstein's theory with more general theories. In particular, it has been found in tensor-scalar theories \cite{19} that a self-gravity as strong as that of a neutron star can naturally (i.e. without fine tuning of parameters) induce order-unity deviations from general relativistic predictions in the orbital dynamics of a binary pulsar thanks to the existence of nonperturbative strong-field effects. [The adjective ``nonperturbative'' refers here to the fact that this phenomenon is nonanalytic in the coupling strength of the scalar field, eq. (\ref{eq:08}), which can be as small as wished in the weak-field limit]. As far as we know, this is the first example where large deviations from General Relativity, induced by strong self-gravity effects, occur in a theory which contains only positive energy excitations and whose post-Newtonian limit can be arbitrarily close to that of General Relativity. A comprehensive account of the use of binary pulsars as laboratories for testing strong-field gravity will be found in ref. \cite{20}. Two complementary approaches can be pursued : a phenomenological one (``Parametrized Post-Keplerian'' formalism), or a theory-dependent one \cite{12}, \cite{20}, \cite{22}. The phenomenological analysis of binary pulsar timing data consists in fitting the observed sequence of pulse arrival times to the generic DD timing formula \cite{21} whose functional form has been shown to be common to the whole class of tensor-multi-scalar theories. The least-squares fit between the timing data and the parameter-dependent DD timing formula allows one to measure, besides some ``Keplerian'' parameters (``orbital period'' $P_b$, ``eccentricity'' $e$,$\ldots$), a maximum of eight ``post-Keplerian'' parameters: $k,\gamma ,\dot{P}_b ,r,s,\delta_{\theta} ,\dot e$ and $\dot x$. Here, $k\equiv \dot{\omega} P_b / 2\pi$ is the fractional periastron advance per orbit, $\gamma$ a time dilation parameter (not to be confused with its post-Newtonian namesake), $\dot{P}_b$ the orbital period derivative mentioned above, and $r$ and $s$ the ``range'' and ``shape'' parameters of the gravitational (``Shapiro'') time delay caused by the companion. The important point is that the post-Keplerian parameters can be measured without assuming any specific theory of gravity. Now, each specific relativistic theory of gravity predicts that, for instance, $k,\gamma, \dot{P}_b ,r$ and $s$ (to quote parameters that have been successfully measured from some binary pulsar data) are some theory-dependent functions of the (unknown) masses $m_1 ,m_2$ of the pulsar and its companion. Therefore, in our example, the five simultaneous phenomenological measurements of $k,\gamma ,\dot{P}_b ,r$ and $s$ determine, for each given theory, five corresponding theory-dependent curves in the $m_1 -m_2$ plane (through the 5 equations $k^{\rm measured} = k^{\rm theory} (m_1 ,m_2 )$, etc$\ldots$). This yields three $(3=5-2)$ tests of the specified theory, according to whether the five curves meet at one point in the mass plane, as they should. [In the most general (and optimistic) case, discussed in \cite{20}, one can phenomenologically analyze both timing data and pulse-structure data (pulse shape and polarization) to extract up to nineteen post-Keplerian parameters.] The theoretical significance of these tests depends upon the physics lying behind the post-Keplerian parameters involved in the tests. For instance, as we said above, a test involving $\dot{P}_b$ probes the propagation (and helicity) properties of the gravitational interaction. But a test involving, say, $k,\gamma ,r$ or $s$ probes (as shown by combining the results of \cite{12} and \cite{19}) strong self-gravity effects independently of radiative effects. Besides the phenomenological analysis of binary pulsar data, one can also adopt a theory-dependent methodology \cite{12}, \cite{20}, \cite{22}. The idea here is to work from the start within a certain finite-dimensional ``space of theories'', i.e. within a specific class of gravitational theories labelled by some theory parameters. Then by fitting the raw pulsar data to the predictions of the considered class of theories, one can determine which regions of theory-space are compatible (at say the 90\% confidence level) with the available experimental data. This method can be viewed as a strong-field genera\-lization of the parametrized post-Newtonian formalism \cite{2} used to analyze solar-system experiments. When non-perturbative strong-field effects are absent one can parametrize strong-gravity effects in neutron stars by using an expansion in powers of the ``compactness'' $c_A \equiv -2 \ \partial \ {\rm ln} \ m_A / \partial \ {\rm ln} \ G \sim G \ m_A / c^2 \ R_A$. Ref. \cite{12} has then shown that the observable predictions of generic tensor-multi-scalar theories could be parametrized by a sequence of ``theory parameters'', $\overline{\gamma} \ , \ \overline{\beta} \ , \ \beta_2 \ , \ \beta' \ , \ \beta'' \ , \ \beta_3 \ , \ (\beta \beta') \ldots$ representing deeper and deeper layers of structure of the relativistic gravitational interaction beyond the first-order post-Newtonian level parametrized by $\overline{\gamma}$ and $\overline{\beta}$. When non-perturbative strong-field effects develop, one cannot use the multi-parameter approach just mentioned. A useful alternative approach is then to work within specific, low-dimensional ``mini-spaces of theories''. Of particular interest is the two-dimensional mini-space of tensor-scalar theories defined by the coupling function $A(\varphi) = {\rm exp} \left( \alpha_0 \, \varphi + {1\over 2} \, \beta_0 \, \varphi^2 \right)$. The predictions of this family of theories (parametrized by $\alpha_0$ and $\beta_0$) are analytically described, in weak-field contexts, by the post-Einstein parameter (\ref{eq:09.a}), and can be studied in strong-field contexts by combining analytical and numerical methods \cite{22}. Let us now briefly summarize the current experimental situation. Concerning the first discovered binary pulsar PSR$1913+16$ \cite{15}, it has been possible to measure with accuracy the three post-Keplerian para\-meters $k, \gamma$ and $\dot{P}_b$. From what was said above, these three simultaneous measurements yield {\it one} test of gravitation theories. After subtracting a small ($\sim 10^{-14}$ level in $\dot{P}_b$ !), but significant, perturbing effect caused by the Galaxy \cite{23}, one finds that General Relativity passes this $( k-\gamma -\dot{P}_b )_{1913+16}$ test with complete success at the $10^{-3}$ level. More precisely, one finds \cite{24}, \cite{18} \begin{eqnarray} \left[ \frac{{\dot{P}}_b^{\rm obs} - {\dot{P}}_b^{\rm galactic}}{{\dot{P}}_b^{\rm GR} [k^{\rm obs} ,\gamma^{\rm obs}]}\right]_{1913+16} \!\!\!\!\!\!\!\!&=\!\!\!\!\!& 1.0032 \pm 0.0023 ({\rm obs})\nonumber \\ \!\!\!\!\!\!\!\!&\pm \!\!\!\!\!& 0.0026 ({\rm galactic}) \nonumber \\ \!\!\!\!\!\!\!\!&=\!\!\!\!\!& 1.0032 \pm 0.0035 \, , \label{eq:00g} \end{eqnarray} where ${\dot{P}}_b^{\rm GR} [k^{\rm obs} ,\gamma^{\rm obs}]$ is the GR prediction for the orbital period decay computed from the observed values of the other two post-Keplerian parameters $k$ and $\gamma$. This beautiful confirmation of General Relativity is an embarrassment of riches in that it probes, at the same time, the propagation {\it and} strong-field properties of relativistic gravity ! If the timing accuracy of PSR$1913+16$ could improve by a significant factor two more post-Keplerian parameters ($r$ and $s$) would become measurable and would allow one to probe separately the propagation and strong-field aspects \cite{24}. Fortunately, the discovery of the binary pulsar PSR$1534+12$ \cite{25} (which is significantly stronger than PSR$1913+16$ and has a more favourably oriented orbit) has opened a new testing ground, in which it has been possible to probe strong-field gravity independently of radiative effects. A phenomenological analysis of the timing data of PSR$1534+12$ has allowed one to measure the four post-Keplerian parameters $k,\gamma ,r$ and $s$ \cite{24}. From what was said above, these four simultaneous measurements yield {\it two} tests of strong-field gravity, without mixing of radiative effects. General Relativity is found to pass these tests with complete success within the measurement accuracy \cite{24}, \cite{18}. The most precise of these new, pure strong-field tests is the one obtained by combining the measurements of $k$, $\gamma$ and $s$. Using the most recent data \cite{stairs} one finds agreement at the 1\% level: \begin{equation} \left[ \frac{s^{\rm obs}}{s^{\rm GR} [k^{\rm obs} ,\gamma^{\rm obs}]}\right]_{1534+12} = 1.007 \pm 0.008 \, . \label{eq:00h} \end{equation} Recently, it has been possible to extract also the ``radiative'' parameter $\dot{P}_b$ from the timing data of PSR$1534+12$. Again, General Relativity is found to be fully consistent (at the $\sim 15\%$ level) with the additional test provided by the $\dot{P}_b$ measurement \cite{stairs}. Note that this gives our second direct experimental confirmation that the gravitational interaction propagates as predicted by Einstein's theory. More recently, measurements of the pulse shape of PSR $1913+16$ \cite{kramer}, \cite{taylorweisberg} have detected a time variation of the pulse shape compatible with the prediction \cite{dr74}, \cite{boc} that the general relativistic spin-orbit coupling should cause a secular change in the orientation of the pulsar beam with respect to the line of sight (``geodetic precession''). As envisaged long ago \cite{dr74} this precession will cause the pulsar to disappear (around 2035) and to remain invisible for hundreds of years \cite{kramer}, \cite{taylorweisberg}. A theory-dependent analysis of the published pulsar data on PSRs $1913+16$, $1534+12$ and $0655+64$ (a dissymetric system constraining the existence of dipolar radiation \cite{WZ89}) has been recently performed within the $(\alpha_0 , \beta_0)$-space of tensor-scalar theories introduced above \cite{22}. This analysis proves that binary-pulsar data exclude large regions of theory-space which are compatible with solar-system experiments. This is illustrated in Fig. 9 of Ref.~\cite{22} which shows that $\beta_0$ must be larger than about $-5$, while any value of $\beta_0$ is compatible with weak-field tests as long as $\alpha_0$ is small enough. \section{Was Einstein 100\% right ?} Summarizing the experimental evidence discussed above, we can say that Einstein's postulate of a pure metric coupling between matter and gravity (``equivalence principle'') appears to be, at least, $99.999 \ \! 999 \ \! 999 \ \! 9\%$ right (because of universality-of-free-fall experiments), while Einstein's postulate (\ref{eq:01}) for the field content and dynamics of the gravitational field appears to be, at least, $99.9\%$ correct both in the quasi-static-weak-field limit appropriate to solar-system experiments, and in the radiative-strong-field regime explored by binary pulsar experiments. Should one apply Occam's razor and decide that Einstein must have been $100\%$ right, and then stop testing General Relativity ? My answer is definitely, no ! First, one should continue testing a basic physical theory such as Ge\-neral Relativity to the utmost precision available simply because it is one of the essential pillars of the framework of physics. This is the fundamental justification of an experiment such as Gravity Probe B (the Stanford gyroscope experiment), which will advance by one order of magnitude our experimental knowledge of post-Newtonian gravity. Second, some very crucial qualitative features of General Relativity have not yet been verified : in particular the existence of black holes, and the direct detection on Earth of gravitational waves. Hopefully, the LIGO/VIRGO network of interferometric detectors will observe gravitational waves early in the next century. Last, some theoretical findings suggest that the current level of precision of the experimental tests of gravity might be naturally (i.e. without fine tu\-ning of parameters) compatible with Einstein being actually only 50\% right ! By this we mean that the correct theory of gravity could involve, on the same fundamental level as the Einsteinian tensor field $g_{\mu \nu}^*$, a massless scalar field $\varphi$. Let us first question the traditional paradigm \cite{10}, \cite{2} according to which special attention should be given to tensor-scalar theories respecting the equivalence principle. This class of theories was, in fact, introduced in a purely {\it ad hoc} way so as to prevent too violent a contradiction with experiment. However, it is important to notice that the scalar couplings which arise naturally in theories unifying gravity with the other interactions systematically violate the equivalence principle. This is true both in Kaluza-Klein theories (which were the starting point of Jordan's theory) and in string theories. In particular, it is striking that (as first noted by Scherk and Schwarz \cite{SS74}) the dilaton field $\Phi$, which plays an essential role in string theory, appears as a necessary partner of the graviton field $g_{\mu \nu}$ in all string models. Let us recall that $g_s = e^{\Phi}$ is the basic string coupling constant (measuring the weight of successive string loop contributions) which determines, together with other scalar fields (the moduli), the values of all the coupling constants of the low-energy world. This means, for instance, that the fine-structure constant $\alpha_{\rm em}$ is a function of $\Phi$ (and possibly of other moduli fields). In intuitive terms, while Einstein proposed a framework where geometry and gravitation were united as a dynamical field $g_{\mu \nu} (x)$, i.e. a soft structure influenced by the presence of matter, string theory extends this idea by proposing a framework where geometry, gravitation, gauge couplings and gravitational couplings all become soft structures described by interrelated dynamical fields. Symbolically, one has $g_{\mu \nu} (x) \sim g^2 (x) \sim G(x)$. This spatiotemporal variability of coupling constants entails a clear violation of the equivalence principle. In particular, $\alpha_{\rm em}$ would be expected to vary on the Hubble time scale (in contradiction with the limit (\ref{eq:05}) above), and materials of different compositions would be expected to fall with different accelerations (in contradiction with the limits (\ref{eq:00.a}), (\ref{eq:00.b}) above). The most popular idea for reconciling gravitational experiments with the existence, at a fundamental level, of scalar partners of $g_{\mu \nu}$ is to assume that all these scalar fields (which are massless before supersymmetry breaking) will acquire a mass after supersymmetry breaking. Typically one expects this mass $m$ to be in the TeV range \cite{CCQR}. This would ensure that scalar exchange brings only negligible, exponentially small corrections $\propto \ \exp (-mr/\hbar c)$ to the general relativistic predictions concerning low-energy gravitational effects. However, the interesting possibility exists that the mass $m$ be in the milli eV range, corresponding to observable deviations from usual gravity below one millimeter \cite{ia}, \cite{fkz}, \cite{dimo}. But, the idea of endowing the scalar partners of $g_{\mu \nu}$ with a non zero mass is fraught with many cosmological difficulties \cite{BS93}, \cite{29}, \cite{DV96b}. Though these cosmological difficulties might be solved by a combination of ad hoc solutions (e.g. introducing a secondary stage of inflation to dilute previously produced dilatons \cite{RT95}, \cite{LS95}), a more radical solution to the problem of reconciling the existence of the dilaton (or any moduli field) with experimental tests and cosmological data has been proposed \cite{30} (see also \cite{28} which considered an equivalence-principle-respecting scalar field). The main idea of Ref. \cite{30} is that string-loop effects (i.e. corrections depending upon $g_s = e^{\Phi}$ induced by worldsheets of arbitrary genus in intermediate string states) may modify the low-energy, Kaluza-Klein type matter couplings $(\propto \, e^{-2 \Phi} \, F_{\mu \nu} \, F^{\mu \nu})$ of the dilaton (or moduli) in such a manner that the VEV of $\Phi$ be cosmologically driven toward a finite value $\Phi_m$ where it decouples from matter. For such a ``least coupling principle'' to hold, the loop-modified coupling functions of the dilaton, $B_i (\Phi) = e^{-2\Phi} + c_0 +c_1 \, e^{2\Phi} +\cdots +$ (nonperturbative terms), must exhibit extrema for finite values of $\Phi$, and these extrema must have certain universality properties. A natural way in which the required conditions could be satisfied is through the existence of a discrete symmetry in scalar space. [For instance, a symmetry under $\Phi \rightarrow -\Phi$ would guarantee that all the scalar coupling functions reach an extremum at the self-dual point $\Phi_m =0$]. A study of the efficiency of this mechanism of cosmological attraction of $\varphi$ towards $\varphi_m$ ($\varphi$ denoting the canonically normalized scalar field in the Einstein frame, see Eq.~(\ref{eq:07})) estimates that the present vacuum expectation value $\varphi_0$ of the scalar field would differ (in a rms sense) from $\varphi_m$ by \begin{equation} \varphi_0 - \varphi_m \sim 2.75 \times 10^{-9} \times \kappa^{-3} \, \Omega_m^{-3/4} \, \Delta \varphi \, . \label{eq:12} \end{equation} Here $\kappa$ denotes the curvature of the gauge coupling function ${\rm ln} \ B_F (\varphi)$ around the maximum $\varphi_m$, $\Omega_m$ denotes the present cosmological matter density in units of $10^{-29} \, g$ cm$^{-3}$, and $\Delta \varphi$ the deviation $\varphi - \varphi_m$ at the beginning of the (classical) radiation era. Equation (\ref{eq:12}) predicts (when $\Delta \varphi$ is of order unity\footnote{However, $\Delta \varphi$ could be $\ll 1$ if the attractor mechanism already applies during an early stage of potential-driven inflation \cite{31}.}) the existence, at the present cosmological epoch, of many small, but not unmeasurably small, deviations from General Relativity proportional to the {\it square} of $\varphi_0 -\varphi_m$. This provides a new incentive for trying to improve by several orders of magnitude the various experimental tests of Einstein's equivalence principle. The most sensitive way to look for a small residual violation of the equivalence principle is to perform improved tests of the universality of free fall. The mechanism of Ref. \cite{30} suggests a specific composition-dependence of the residual differential acceleration of free fall and estimates that a non-zero signal could exist at the very small level \begin{equation} \left( {\Delta a \over a} \right)_{\rm rms}^{\rm max} \sim 1.36 \times 10^{-18} \, \kappa^{-4} \, \Omega_m^{-3/2} \, (\Delta \varphi)^2 , \label{eq:13} \end{equation} where $\kappa$ is expected to be of order unity (or smaller, leading to a larger signal, in the case where $\varphi$ is a modulus rather than the dilaton). Let us emphasize that the strength of the cosmological scenario considered here as counterargument to applying Occam's razor lies in the fact that the very small number on the right-hand side of eq. (\ref{eq:13}) has been derived without any fine tuning or use of small parameters, and turns out to be naturally smaller than the $10^{-12}$ level presently tested by equivalence-principle experiments (see equations (\ref{eq:00.a}), (\ref{eq:00.b})). The estimate (\ref{eq:13}) gives added significance to the project of a Satellite Test of the Equivalence Principle (nicknamed STEP, and currently studied by NASA, ESA and CNES) which aims at probing the universality of free fall of pairs of test masses orbiting the Earth at the $10^{-18}$ level \cite{32}.
\section{Introduction} A major discovery from ASCA has been the discovery of a clear, broad, skewed iron line in the spectrum of the Seyfert galaxy MCG--6-30-15 (Tanaka et al 1995). This emission line has a profile matching that expected from the inner regions, from about 6 to 40 gravitational radii (i.e. 6 -- 40 $GM/c^2$), of a disk around a black hole (Fabian et al 1989). No simple alternative model is capable of explaining this profile (Fabian et al 1995). Similar skewed lines have since been found in the spectra of many other Seyfert galaxies (Nandra et al 1997; Reynolds 1997). The broad line in MCG--6-30-15 has also been clearly detected with BeppoSAX (Guainazzi et al 1999). The 1994 ASCA observation of MCG--6-30-15 reported by Tanaka et al (1995) remains however the best example of a broad line due to the good spectral resolution of the detectors used and the long integration time of 4.5 days. Here we report on a similar long ASCA observation of the object made in 1997. We confirm in detail the time-averaged line shape, which only shows a small change in the `blue' horn. During the 1994 ASCA observation the light curve of the source showed both a flare and a deep minimum (Iwasawa et al 1996). The line profile was seen to alter, being mostly a blue horn during the flare and then showing only an extreme red horn during the minimum. These changes were assumed to be due to changes in the location in the most active regions irradiating the disk (and so producing the iron line), the flare being on the approaching side of the disk and the minimum emission from within the innermost stable orbit of a non-spinning Schwarzschild black hole (Iwasawa et al 1996). This last possibility has been explored further by Dabrowski et al (1997), Reynolds \& Begelman (1998); Weaver \& Yaqoob (1998) and Young, Ross \& Fabian (1998). The light curve of the source during 1997 also shows flares and dips. We examine in detail here the major flare seen during which both the continuum and line show large changes. \section{Observations and data reduction} MCG--6-30-15 was observed with ASCA from 1997 August 3 to 1997 August 10 with a half-day gap in the middle. It was also observed simultaneously with Rossi X-ray Timing Explorer (RXTE) (Lee et al 1999). The Solid state Imaging Spectrometer (SIS; S0 and S1) was operated in Faint mode throughout the observation, using the standard CCD chips (S0C1 and S1C3). The Gas Imaging Spectrometer (GIS; G2 and G3) was operated in PH mode. We present results mainly from the SIS data in this Letter. The ASCA S0 light curve in the 0.6--10 keV band is shown in Fig. 1. \begin{figure} \centerline{\psfig{figure=fig1.ps,width=0.45\textwidth,angle=270}} \caption{The ASCA SIS0 0.6--10 keV light curve of MCG--6-30-15 in August 1997. The epoch of the light curve is 1997 August 3, 22:06:11 (UT). Each data point has a 128-s (or less) exposure time.} \end{figure} Data reduction was carried out using FTOOLS version 4.0 and 4.1 with standard calibration provided by the ASCA Guest Observer Facility (GOF). The good exposure time is approximately 231 ks from each SIS detector. The source counts are collected from a region centred at the X-ray peak within 4 arcmin in radius for the SIS and 5 arcmin for the GIS. The background data are taken from a (nearly) source-free region in the same detector with the same observing time. The efficiency of the S1 detector below 1 keV appears to be severely reduced due to the Residual Darkframe Distribution (RDD, Dotani 1998), which the current response matrix (generated from calibratione files in the FTOOLS version 4.1 release) does not take into account for. The RDD effect on the S0 data from 1CCD observations has been found to be very little (Dotani 1998). Therefore, the S1 data below 1 keV were discarded for the spectral analysis presented here. The energy resolution of the SIS at 6.4 keV when the observation was carried out had degraded to $\sim 250$ eV (FWHM), about twice that attained immediately after launch of the satellite. \section{Comparisons with the 1994 long observation} The average count rates in the 0.6--10 keV band from the S0/S1 detectors are 1.16/0.93 ct\thinspace s$^{-1}$ (cf. 1.53/1.25 ct\thinspace s$^{-1}$\ during the previous long observation in 1994). The average observed fluxes are $1.47\times 10^{-11}$erg~cm$^{-2}$~s$^{-1}$\ in the 0.5--2 keV band and $3.41\times 10^{-11}$erg~cm$^{-2}$~s$^{-1}$\ in the 2--10 keV band. \subsection{Total energy spectrum and warm absorber} The observed 0.6--10 keV X-ray flux during the present observation is lower by 26 per cent than that during the 94 long observation. The 3--10 keV (the iron K band, 4--7.5 keV, excluded) power-law slope is $\Gamma = 1.94^{+0.06}_{-0.07}$, which is similar to the 94 data. Features of the warm absorber detected are two edges due to OVII at 0.72 keV and OVIII at 0.85 keV (e.g., Otani et al 1996) and one at 1.1 keV, probably due to NeIX and/or Fe L. The 97 spectrum is harder in the low energy band than the 94 spectrum, which may be explained by an increase in absorption. Details will be reported by Matsumoto et al (in prep). \subsection{The iron K line} The 3--10 keV data were investigated for iron K line emission. The continuum spectrum was modelled with a power-law reflection model ({\tt pexrav}, Magdziarz \& Zdziarski 1995) modified by cold absorption of $N_{\rm H}$\thinspace $= 7\times 10^{20}$cm$^{-2}$ (which has virtually no effect on the 3--10 keV continuum). The parameters of {\tt pexrav}, apart from photon index and normalization, were matched to the previous measurements of MCG--6-30-15: the cut-off energy, 130 keV (from the BeppoSAX observation by Guainazzi et al 1999); reflection intensity, corresponding to $\Omega/2\pi =1$ (Guainazzi et al 1999; Lee et al 1999); iron abundance of unity (Lee et al 1999); and inclination of the reflecting slab, 30$^{\circ}$ (Tanaka et al 1995). The line feature is fitted by the diskline model for a Schwarzschild black hole (Fabian et al 1989). The rest energy of the line emission is assumed to be 6.4 keV, appropriate for cold iron. A power-law ($\Gamma = 1.96^{+0.04}_{-0.03}$) modified by reflection plus a diskline model provide a good fit ($\chi^2 = 724.7$ for 729 degrees of freedom). The best-fit parameters of the diskline model are shown in Table 1. \begin{figure} \centerline{\psfig{figure=fig2.ps,width=0.4\textwidth,angle=270}} \caption{The iron K emission-line profiles of MCG--6-30-15 obtained from the long observations in 1997 (upper panel) and 1994 (lower panel from Tanaka et al 1995). Dotted line shows best-fit diskline model for a Schwarzschild black hole. See Table 1 for the diskline parameters for the 97 profile.} \end{figure} \begin{table*} \begin{center} \caption{The best-fit parameters of the diskline fit to the averaged Fe K line profile. The continuum is modelled by a power-law modified by reflection ({\tt pexrav}, see text). The diskline model by Fabian et al (1989) is used. (1) The rest energy of the line emission ; (2) power-law index of the radial emissivity law ($\propto r^{-\alpha}$); (3),(4)Inner and outer radii of the line emitting disk in a unit of gravitational radius ($r_{\rm g}$); (5) inclination angle of the accretion disk; (6) efficiency-corrected line intensity; (7) equivalent width of the line emission calculated assuming all the line emission is concentrated in a narrow energy range around 6.4 keV.} \begin{tabular}{ccccccc} (1) & (2) & (3) & (4) & (5) & (6) & (7) \\ $E$ & $\alpha $ & $R_{\rm in}$ & $R_{\rm out}$ & $i$ & $I$ & $EW$ \\ keV & & $r_{\rm g}$ & $r_{\rm g}$ & deg & ph\thinspace s$^{-1}$\thinspace cm$^{-2}$ & eV \\[5pt] 6.4 & $4.1^{+2.0}_{-2.0}$ & $6.7^{+0.9}_{-0.7}$ & $24_{-10}^{+20}$ & $32^{+2}_{-2}$ & $1.36^{+0.26}_{-0.24}\times 10^{-4}$ & $388^{+74}_{-68}$ \\ \end{tabular} \end{center} \end{table*} The efficiency-corrected line profile\footnote{This is not an `unfolded' spectrum but obtained from the ratio, the data divided by the power-law model (folded through the detector response) best-fitting the neighbouring continuum, multiplied by the power-law (in original form). The plot is therefore independent from the model used for fitting the line.} for the present data set is shown in Fig. 2, along with the one from the previous long observation in 1994 (Tanaka et al 1995). The profile from the present observation appears to be less bright in the blue peak while it shows a slightly more extended red wing. During the previous observation, the bright flare ({\it i-3}) data showed a narrow-core-dominated line profile (Iwasawa et al 1996). Such a line shape is not found during the present observation. However, this is not sufficient to explain the difference in the blue horn of the time-averaged line profiles between the two observations. Although the overall line shape is similar between 94 and 97, the steeper radial emissivity index suggests that the mean weight of the line emissivity may be slightly shifted towards the inner part of the accretion disk in 97 as compared to 94. \section{The major flare} Changes in the iron line profile were investigated in time sequence, details of which will be reported elsewhere. Here we show the peculiar behaviour of the energy spectrum and iron line during the major flare, which occured around $1.9\times 10^5$ s in the light curve (see Fig. 1 and Fig. 3 for a detailed version of the light curve around the flare). The continuum is steeper than usual ($\Delta\Gamma \sim 0.17$), particularly at low energies (1--3 keV band), as shown in Fig. 4. \begin{figure} \centerline{\psfig{figure=fig3.ps,width=0.4\textwidth,angle=270}} \caption{The ASCA SIS0 light curve of MCG--6-30-15 around the brightest flare. The epoch of the light curve is 1997 August 3, 22:06:11. The iron K line profiles during the time intereval {\it a} and {\it b} are shown in Fig. 5.} \end{figure} \begin{figure} \centerline{\psfig{figure=fig4.ps,width=0.4\textwidth,angle=270}} \caption{The major flare spectrum (taken from the time-intereval {\it a} in Fig. 3, squares: S0; triangles: S1) divided by the best-fit model for the time-averaged spectrum. The same plot for the time-averaged spectrum is also shown. The model is a power-law ($\Gamma = 1.94$) modified by three absorption edges at 0.71, 0.85, and 1.1 keV, and cold absorption ($N_{\rm H}$\thinspace $=9\times 10^{20}$cm$^{-2}$). The Fe K band was excluded from the fit so that the Fe K line features remain in the plot.} \end{figure} \begin{figure} \centerline{\psfig{figure=fig5.ps,width=0.4\textwidth,angle=0}} \caption{The ratio plot of the SIS data to the baseline power-law models for the time intervals {\it a} and {\it b} in Fig. 3. The vertical dotted line in the upper panel indicates the rest energy of the cold iron K$\alpha$ emission line, 6.4 keV. Almost all of the line emission is redshifted well below the rest line energy during the flare ({\it a}). When the flare ceased ({\it b}), the line shape recovered to the ordinary one, as shown in the lower panel.} \end{figure} \subsection{Iron K line emission with large redshift?} Excess emission above a power-law continuum, which is presumably due to broad iron K line emission, is found in the 4--7 keV range. The ratio plots of the data and the baseline power-law model for the intervals of the flare peak ({\it a}) and of the subsequent dip ({\it b}) are shown in Fig. 5. The baseline model is obtained by fitting a power-law to the data adjacent to the iron line band. The line profile of the flare peak ({\it a}) shows a sharp decline at $\sim $5.6 keV, which is far below the rest energy of the line emission of Fe~K$\alpha$, 6.4 keV, and a red-wing extending down to $\sim $3.5 keV. We have checked the GIS data which confirm the SIS result. No line emission is detected at 6.4 keV ($<3\times 10^{-5}$ph\thinspace s$^{-1}$\thinspace cm$^{-2}$, $EW<60$ eV, 90 per cent upper limits obtained from the joint fit to the SIS and GIS data). This extremely redshifted line profile cannot be explained by the diskline for a Schwarzschild black hole because of insufficient gravitational redshift (it may arise from infalling gas, as proposed by Reynolds \& Begelman 1997, but then there would also be a large absorption edge; Young et al 1998). On the other hand, the diskline model for a Kerr black hole by Laor (1991; which is for a maximally-rotating black hole) gives a good fit. The result of this diskline fit is shown in Table 5, where the rest line energy, the inner radius and inclination of the disk are assumed to be 6.4 keV, 1.235$r_{\rm g}$, and 30$^{\circ}$. The outer radius is constrained well at $(5\pm 1)$$r_{\rm g}$\ due to the well-defined decline of the redshifted, blue peak. The inferred negative emissivity index (it is poorly constrained) and the well-constrained outer radius perhaps suggest the line emission is concentrated in annuli around $\sim$ 5$r_{\rm g}$. A fit with a double-gaussian to the line profile is also given in Table 5. The line intensity of the line is about 3 times larger than that of the time-averaged one. The EW is $\sim 700$ eV when computed with respect to the continuum at 6.4 keV but $\sim 400$ eV to the continuum in the energy range of the observed line. The line profile models available in XSPEC, the fits of which have just been reported above, are for complete disk annuli. It is possible however that just part of an annulus of the disk is irradiated during a flare, i.e. that part immediately below the flare itself. This offers further possible locations for the flare such as on the receding side of the disk where the peak at 5 keV is mostly due to the doppler effect, or on the approaching side much closer in where it is due to gravitational redshift. We show in Fig. 6 the locus of points which cause a 6.4 keV line in the disk frame to appear at 5 keV for an observer seeing the disk around a maximally spinning black hole at an inclination of $30^{\circ}$. We have also fit the flare spectrum with model line profiles created from a disk divided into 36 azimuthal sectors and 24 radial bins between 1 and 25 $r_{\rm g}$. Although acceptable fits at $\sim 3\sigma$ level are found for most of the points around this locus, the region at small radii ($\sim 2.5$$r_{\rm g}$) on the approaching side of the disk is favoured most (see the confidence contours in Fig. 6). This is mainly due to the excess flux around 4~keV shown in Fig~5a and not the choice of grid size. In the subsequent time-interval {\it b}, the averaged 0.6--10 keV count rate dropped by a factor of 2.2, compared with the flare interval (see Fig. 3). The line shape has then recovered to the ordinary one, as seen in the time-averaged spectrum (Fig. 5b). The line intensity, $(2.2\pm 1.3)\times 10^{-4}$ph\thinspace s$^{-1}$\thinspace cm$^{-2}$, also dropped by a factor of 2. \begin{figure} \vspace{-5mm} \centerline{\psfig{figure=fig6a.ps,width=0.4\textwidth}} \vspace{-5mm} \centerline{\psfig{figure=fig6b.ps,width=0.38\textwidth,angle=270}} \caption{Upper panel: Locus of points which shift a 6.4keV line to 5~keV. The accretion disk around a maximally spinning (angular momentum, a = 0.998) black hole viewed at an inclination angle of $30^{\circ}$ is assumed. The approaching side of the disk is on the left (the receding side on the right). Lower panel: Confidence contours in the radius and azimuthal-angle plane obtained on fitting model lines from separate parts of a disk to the line profile data shown in Fig. 5a. Azimuthal angle $\phi$ is measured clock wise from the near side of the disk. Contours are drawn at 68, 90 and 99 per cent confidence levels for two parameters of interest. The possibility that a whole ring at $\sim 5$$r_{\rm g}$\ yields an acceptable fit (see text) is not evident from this plot, since if it involves only single small azimuthal sector at each radius.} \end{figure} \begin{table*} \begin{center} \caption{Diskline and double-gaussian fits to the line profile from the flare interval ({\it a} in Fig. 3). The diskline model is for a maximally rotating Kerr hole by Laor (1991). The emissivity index is defined by a power-law function of radius ($\propto r^{-\alpha}$).} \begin{tabular}{cccccccc} \multicolumn{7}{c}{Diskline (Laor 1991)} \\[5pt] $E$ & $\alpha $ & $R_{\rm in}$ & $R_{\rm out}$ & $i$ & $I$ & $\chi^2$/dof \\ keV & & $r_{\rm g}$ & $r_{\rm g}$ & deg & $10^{-4}$ph\thinspace s$^{-1}$\thinspace cm$^{-2}$ & \\[5pt] 6.4 & $-4.5(<2.7)$ & 1.235 & $4.6^{+0.6}_{-0.4}$ & 30 & $4.9^{+3.0}_{-2.2}$ & 102.0/143 \\[8pt] \multicolumn{7}{c}{Double-gaussian} \\[5pt] $E_1$ & $\sigma_1$ & $I_1$ & $E_2$ & $\sigma_2$ & $I_2$ & $\chi^2$/dof \\ keV & keV & $10^{-4}$ph\thinspace s$^{-1}$\thinspace cm$^{-2}$ & keV & keV & $10^{-4}$ph\thinspace s$^{-1}$\thinspace cm$^{-2}$ & \\[5pt] $5.39^{+0.16}_{-0.23}$ & $0.19(<0.44)$ & $2.0^{+2.6}_{-1.6}$ & $4.49^{+0.57}_{-1.30}$ & $0.51(<1.93)$ & $2.4^{+4.6}_{-2.1}$ & 101.6/140 \\ \end{tabular} \end{center} \end{table*} \section{Discussion} The long ASCA observation of 1997 has confirmed in detail the broad iron line in the Seyfert galaxy MCG--6-30-15. The time-averaged emission appears to originate from a disk extending between about 6 and 40 $r_{\rm g}$\ of a massive black hole. If the rapid variability of this source is due to flares above the accretion disk, then the long term constancy of the line profile indicates that there are usually several flares at once on the disk and that the distribution of flares is almost constant in a time-averaged sense. The spectrum of the source, and in particular the line profile, changed dramatically during a bright flare. Unlike the bright blue horn apparent during the 1994 flare, we now see essentially a bright red horn. If this is interpreted in the context of a relativistic diskline, the dominant flare must occur at smaller radii than usual. There are two possible locations of the flare, depending on the mass of the black hole in MCG--6-30-15. One interesting possibility is a flare localized on the approaching side of the disk at $\sim 2.5$$r_{\rm g}$\ (see Fig. 6). The duration of the flare is about 1 hour while the Keplerian orbital time around a $10^7M_7${\it M}$_\odot$\ black hole is $10^4M_7r_1^{3/2}$s at $10r_1$$r_{\rm g}$. If the flare is confined within, say, 1/6 orbit at 2.5$r_{\rm g}$\ (see Fig. 6), the duration of the flare (and peculiar line shape) requires the black hole mass to be larger than $20 M_7${\it M}$_\odot$. Therefore this solution is valid only if the black hole in MCG--6-30-15 is more massive than $10^8${\it M}$_\odot$. The spectral fit with the model for azimuthally-averaged line emission (Laor 1991) suggested that the line emission may be produced in a narrow range of radii around 5$r_{\rm g}$\ during the major flare (see Table 2). The duration of the flare corresponds to $1M_7^{-1}$ orbital time at 5$r_{\rm g}$. This is the preferred solution if the black hole mass is significantly smaller than $10^8${\it M}$_\odot$. There are some difficulties for the first interpretation. In order to restrict the line production to part of the disk, the flare must be placed very close to the disk surface. A flare on the approaching side of the disk is generally expected to be amplified due to relativistic beaming (e.g., Karas et al 1992), which appears to be consistent with the observed flux variation. However, at a small radii such as 2.5$r_{\rm g}$\ on a disk inclined at $30^{\circ}$, gravitational redshift and frame dragging overwhelm doppler boosting so that the emission reaching a distant observer is suppressed by more than an order of magnitude when the X-ray source is placed at 1$r_{\rm g}$\ above the disk (M. Ruszkowski, priv. comm.). Therefore the flare would have to be intrinsically much more intense than observed. This may be possible if the emitted power increases rapidly towards inner radii around a spinning black hole. The strong light deflection implies that the reflection from the disk should also be enhanced by a factor of $\sim 2$ at the same time (see also Martocchia \& Matt 1996). The closeness of the continuum source may also cause the disk surface to be highly ionized. Although the high energy end of the ASCA data is rather noisy, the 6--10 keV spectrum during the flare ($\Gamma = 1.5\pm 0.4$) suggests a possible spectral flattening which could be due to strong reflection. The interpretations discussed above are, of course, not unique, but both require that the accretion disk extends close to the central black hole and that it spins rapidly, as suggested by Iwasawa et al (1996). Although the profile of the broad line in MCG--6-30-15 appears to be fairly constant in a time-averaged sense, it does undergo dramatic changes every few days. Such changes offer interesting possibilities with which to probe different parts of the disk and to map the innermost regions about the black hole. The lack of a narrow 6.4 keV line during the flare (the 90 per cent upper limit of intensity is only 40 per cent of the blue peak intensity of the time-averaged line) confirms the suggestion made by the previous ASCA observation (Iwasawa et al 1996) that there is little line emission from far out in the disk or torus. A narrow 6.4 keV line might be delayed by an hour or so, if it is produced around $100M_7$$r_{\rm g}$. Any line emission from farther out should be more constant because the variability of the line is smeared out. Evidence for such line emission appears to be weak. \section*{Acknowledgements} We thank all the members of the ASCA team. ACF and KI thank Royal Society and PPARC, respectively, for support. Chris Reynolds is thanked for his useful comments.
\section*{Introduction} For a long time, it was assumed that neutrinos have vanishing quantities: $m_\nu=0, Q_{\rm em}=0$, and $\mu_\nu=0$. Among these the neutrino mass problem has attracted the most attention, and finally we might have an evidence for nonzero neutrino mass~\cite{Super}. The other important neutrino property to be exploited is the electromagnetic property, in particular the magnetc moment. The reason for vanishing neutrino mass was very naive in 50's and 60's: the hypothesis of the $\gamma_5$-invariance. Under the $\gamma_5$-invariance, $\nu=\pm \gamma_5\nu$, $\nu$ appears only in one chirality. In the standard model (SM), this is encoded as no right-handed neutrino. In gauge theory models, the story changes because one can calculate the properties of the neutrino at high precision. In SM, one cannot write a mass term for $\nu$ in $d\le 4$ terms. To write a mass term for $\nu$, one has to introduce $d\ge 5$ terms, or introduce singlet neutrino(s). The two-component neutrino we consider in the left-handed doublet can be Weyl or Majorana type. If it is a Weyl neutrino, it satisfies $\nu_i=a_i\gamma_5\nu_i$ where $a_i=1$ or $-1$. Then the magnetic moment term is given by $\bar\nu_i\sigma_{\mu\nu}q^\nu\nu_j-\bar\nu_j\sigma_{\mu\nu}q^\nu\nu_i \rightarrow -a_ia_j[\bar\nu_i\sigma_{\mu\nu}q^\nu\nu_j-\bar \nu_j\sigma_{\mu\nu}q^\nu\nu_i]. $ Therefore, to have a nonvanishing magnetic moment (or mass), we must require $a_ia_j=-1$, i.e. the existence of right-handed singlet neutrino(s). For Majorana neutrinos, $\psi_c=C\psi^*$, it is possible to write a mass term without introducing right-handed neutrino(s). Thus, it is possible to introduce neutrino masses and magnetic moments in the SM, by a slight extension of the model. The question is how large they are. For the detection of neutrino masses and oscillations, there have been numerous studies from solar-, atmospheric-, reactor-, and accelerator-neutrino experiments. The effect of neutrinos in cosmology was also used to get bounds on neutrino masses. On the other hand, for the neutrino magnetic moment astrophysical constraints gave useful bounds. Usually, the bound of the neutrino magnetic moment is given in units ($f$) of Bohr magneton ($\mu_B$), \begin{equation} \mu_{\nu_i}=f_i\mu_B. \end{equation} \section*{History and the known bounds} The first significant bound on magnetic moment of $\nu_e$ was given from astrophysics, $|f_e|<10^{-10}$, by Bernstein et al.~\cite{bernstein}. A better bound on $f_e$ was obtained from SN1987A, $|f_e|<10^{-13}$~\cite{SN}. For the muon neutrino, the useful bound was obtained from the neutral current data~\cite{kmo}, $|f_\mu|<0.8\times 10^{-8}$. The first bound on the transition magnetic moment was given also from the neutral current experiment~\cite{kim}. For the tau neutrino, $|f_\tau|<1.3\times 10^{-7}$ has been obtained recently~\cite{sergei}. For the theoretical side, it has been known from early days that it is possible to generate large magnetic moments for neutrinos~\cite{cal}. Note that the see-saw mass for neutrinos appear as in Fig.~1. Here, there does not exist any charged particle and hence there is no contribution to magnetic moment at this level. Thus to have a large neutrino magnetic moments, one needs a Feynman diagram of Fig.~2 type, where we introduced a heavy lepton $L$ coupling to $W$ via \begin{figure}[b!] \centerline{\epsfig{file=nmm1.eps,height=4cm, width=6.67cm}} \vspace{10pt} \caption{A see-saw mechanism for neutrino masses.} \label{fig1} \end{figure} \begin{equation} \left(\matrix{\nu_l\cr l^-\cos\alpha+L^-\sin\alpha}\right)_L\ ,\ \left(\matrix{L^0\cr -l^-\sin\alpha+L^-\cos\alpha}\right)_L\ ,\ \left(\matrix{\nu_l\cr L^-}\right)\ ,\ \left(\matrix{L^0\cr l^-}\right). \end{equation} \begin{figure}[b!] \centerline{\epsfig{file=nmm2.eps,height=4cm, width=6.67cm}} \vspace{10pt} \caption{A Feynman diagram for neutrino magnetic moment.} \label{fig2} \end{figure} Then one can easily estimate the magnetic moment of neutrinos as~\cite{cal}\footnote{$f^\prime$ is the transition moment.} \begin{equation} f\ {\rm or}\ f^\prime\ =\ \frac{G_Fm_Lm_e}{2\sqrt{2}\pi^2}ab I \left(\frac{1}{2}+\frac{1}{2}\delta_{\nu\nu^\prime}\right) \end{equation} where $abI$, which is a function of mixing and Feynman integral, is of order 1. One can also draw Feynman diagrams with charged scalars in the loop with appropriate Yukawa interactions introduced. This kind of diagrams generally introduce transition magnetic moment of order$ f^\prime \sim {m_Lm_e}/{M^2}$ where $M$ is the mass of the intermediate scalar or gauge boson. Note that without extra charged leptons $m_L$ should be neutrino mass, rendering an extremely small $\mu_\nu$. \section*{NC, $\mu^\prime_\nu$, and Single $\pi^0$ Production by $\nu_{\mu}^{\rm atm}$} The Super-K collaboration has reported the ratio \begin{equation} R_{\pi^0/e}=\frac{(\pi^0/e)_{\rm data}}{(\pi^0/e)_{\rm MC}} =0.93\pm 0.07\pm 0.19 \end{equation} which is consistent with 1 at present. However, one may narrow down the experimental errors and can observe whether it is different from 1 or not. One assumes that $\nu_e$ is not oscillated in the atmospheric neutrino data sample, and hence the MC electrons are estimated with the standard CC cross section. The denominator is calculated assuming that the NC is the same for the cases with and without neutrino oscillation. So $R_{\pi^0/e}$ is expected to be 1 if there is no oscillation of SM neutrinos to sterile neutrinos. If there exist oscillation of $\nu_\mu$ to sterile neutrinos, then one expects that $R_{\pi^0/e}<1$. However, in our recent work~\cite{kkl} we pointed out that one should be careful to draw a firm conclusion on this matter because if a sizable transition magnetic moment of $\nu_\mu$ exists then one expects a different conclusion. For the study of NC, the single $\pi^0$ production is known to be very useful. Most dominant contribution to the single $\pi^0$ production at the atmospheric neutrino energy is through $\Delta$ production, \begin{equation} \nu+N\rightarrow\nu+\Delta\ ,\ \ \Delta\rightarrow N+\pi^0. \end{equation} In this calculation, we used the form factors given in Ref.~\cite{fogli}. For $E_\nu<10$~GeV or the kinetic energy of recoil nucleon $<1$~GeV, the process $\nu+N\rightarrow \nu^\prime +N$ is difficult to observe at Super-K. So the $\pi^0$ production is the cleanest way to detect NC interactions through Cherenkov ring (from $\pi^0$ decay) at Super-K. For transition magnetic moment parametrized by $f^\prime$, $ if^\prime\mu_B\bar u(l^\prime)\sigma_{\mu\nu}q^\nu u(l)_{\nu_\mu} $ where $q=l-l^\prime$, the single $\pi^0$ production cross section through $\Delta$ production is given in Ref.~\cite{kkl}. In Fig.~3, we show the result (the ratio of the neutrino magnetic moment contribution and the NC contribution) as a function of neutrino energy. \begin{figure}[b!] \centerline{\epsfig{file=nmm3.eps,height=4cm, width=6.67cm}} \vspace{10pt} \caption{The ratio of $\mu^\prime_{\nu_\mu}$ and NC contributions as a function of $E_{\nu_\mu}$. See Ref.~[7] for details.} \label{fig3} \end{figure} Note that the magnetic moment part is more important at low $q^2$ region due to the photon propagator. In principle, one can distinguish neutrino magnetic moment interactions from the NC interactions. From Fig.~3, if we require $r_{f^\prime/NC}\le 0.13$, then we obtain a bound $f^\prime \le 2.2\times 10^{-9}$. In conclusion, the transition magnetic moment $f^\prime$ can be large. For $\nu^\prime$ heavy, it is not restricted by SN1987A bound. But atmospheric neutrinos of 1--10~GeV can produce $\nu^\prime$, and can mimick NC data~\cite{kim}. Before interpreting NC effects from atmospheric neutrino data, one has to separate out the $\mu_{\nu}^\prime$ contribution. \section*{Models with Large $\mu_\nu$} Before closing, we point out $\mu_\nu$ and solar neutrino problem. One possibility to reduce solar $\nu_e$ flux is to precess $\nu_{eL}$ to $\nu_R$ with a large $\mu_\nu$ in a strong magnetic field~\cite{okun}. But this idea seems to be ruled out by the nucleosynthesis argument~\cite{morgan}, $\mu_\nu<10^{-11}\mu_B$, and the SN1987A argument~\cite{SN}, $\mu_\nu<10^{-13}\mu_B$. The SN1987A bound is coming from the energy loss mechanism: if $\nu_{eR}$ is created, it takes out energy out of the core. But if it is trapped, then the bound does not apply.\footnote{The transition magnetic moments to $\nu^\prime$ are not restricted by these bounds for a heavy enough $\nu^\prime$, but then the transition magnetic moment cannot account for the solar neutrino deficit.} The solar neutrino problem requires $\mu_{\nu_e}\sim 10^{-11}\mu_B$ which is considered to be large. To use the idea of trapping, the oscillation is \begin{equation} \nu_{eL}\rightarrow \nu^c_{\mu R}. \end{equation} Namely, we use the Konopinski-Mahmoud scheme where $\nu^c_{\mu R}$ is the weak interaction partner of $\mu^+_R$, implying the oscillated neutrino participates in weak interactions; hence trapped in the supernova core. This picture is very restrictive as suggested by models given in Ref.~\cite{voloshin}. These models try to get a large neutrino magnetic moment while keeping the neutrino mass small. The $SU(2)_V$ symmetry in which $(\nu_e, \nu^c_{\mu R})$ forms a doublet under $SU(2)_V$ is introduced for this purpose. In this case, $\mu$-number minus electron-number is conserved. The reason for vanishing mass is that $\nu^T C\nu^c$ is symmetric under $SU(2)_V$, i.e. the mass term is a triplet under $SU(2)_V$ and hence is forbidden. On the other hand the magnetic moment term, $\nu^T C\sigma_{\alpha\beta}\nu^c F^{\alpha\beta}$, is a singlet and is allowed, as given in Fig.~4. But, {\it why $SU(2)_V$?} It is like the dilemma asked in any fermion mass ansatz problem. \begin{figure}[b!] \centerline{\epsfig{file=nmm4.eps,height=4cm, width=6.67cm}} \vspace{10pt} \caption{Neutrino magnetic moment with $SU(2)_V$.} \label{fig4} \end{figure}
\section{INTRODUCTION} This is the sixth part of our eight presentations in which we consider applications of methods from wavelet analysis to nonlinear accelerator physics problems. This is a continuation of our results from [1]-[8], in which we considered the applications of a number of analytical methods from nonlinear (local) Fourier analysis, or wavelet analysis, to nonlinear accelerator physics problems both general and with additional structures (Hamiltonian, symplectic or quasicomplex), chaotic, quasiclassical, quantum. Wavelet analysis is a relatively novel set of mathematical methods, which gives us a possibility to work with well-localized bases in functional spaces and with the general type of operators (differential, integral, pseudodifferential) in such bases. In contrast with parts 1--4 in parts 5--8 we try to take into account before using power analytical approaches underlying algebraical, geometrical, topological structures related to kinematical, dynamical and hidden symmetry of physical problems. In part 2 according to the orbit method and by using construction from the geometric quantization theory we construct the symplectic and Poisson structures associated with generalized wavelets by using metaplectic structure. In part 3 we consider applications of very useful fast wavelet transform technique (FWT) (part 4) to calculations in quasiclassical evolution dynamics. This method gives maximally sparse representation of (differential) operator that allows us to take into account contribution from each level of resolution. \section{Metaplectic Group and Representations} Let $Sp(n)$ be symplectic group, $Mp(n)$ be its unique two- fold covering -- metaplectic group [9]. Let V be a symplectic vector space with symplectic form ( , ), then $R\oplus V$ is nilpotent Lie algebra - Heisenberg algebra: $[R,V]=0, \ [v,w]=(v,w)\in R,\ [V,V]=R$. $Sp(V)$ is a group of automorphisms of Heisenberg algebra. Let N be a group with Lie algebra $R\oplus V$, i.e. Heisenberg group. By Stone-- von Neumann theorem Heisenberg group has unique irreducible unitary representation in which $1\mapsto i$. Let us also consider the projective representation of symplectic group $Sp(V)$: $U_{g_1}U_{g_2}=c(g_1,g_2)\cdot U_{g_1g_2}$, where c is a map: $Sp(V)\times Sp(V)\rightarrow S^1$, i.e. c is $S^1$-cocycle. But this representation is unitary representation of universal covering, i.e. metaplectic group $Mp(V)$. We give this representation without Stone-von Neumann theorem.\ Consider a new group $F=N'\bowtie Mp(V),\quad \bowtie$ is semidirect product (we consider instead of $ N=R\oplus V$ the $ N'=S^1\times V, \quad S^1=(R/2\pi Z)$). Let $V^*$ be dual to V, $G(V^*)$ be automorphism group of $V^*$.Then F is subgroup of $ G(V^*)$, which consists of elements, which acts on $V^*$ by affine transformations. This is the key point! Let $q_1,...,q_n;p_1,...,p_n$ be symplectic basis in V, $\alpha=pdq=\sum p_{i}dq_i $ and $d\alpha$ be symplectic form on $V^*$. Let M be fixed affine polarization, then for $a\in F$ the map $a\mapsto \Theta_a$ gives unitary representation of G: $ \Theta_a: H(M) \rightarrow H(M)$. Explicitly we have for representation of N on H(M): $ (\Theta_qf)^*(x)=e^{-iqx}f(x), \ \Theta_{p}f(x)=f(x-p) $ The representation of N on H(M) is irreducible. Let $A_q,A_p$ be infinitesimal operators of this representation $$ A_q=\lim_{t\rightarrow 0} \frac{1}{t}[\Theta_{-tq}-I], \quad A_p=\lim_{t\rightarrow 0} \frac{1}{t}[\Theta_{-tp}-I], $$ $$\mbox{then } A_q f(x)=i(qx)f(x),\ A_p f(x)=\sum p_j\frac{\partial f}{\partial x_j}(x) $$ Now we give the representation of infinitesimal ba\-sic elements. Lie algebra of the group F is the algebra of all (non\-ho\-mo\-ge\-ne\-ous) quadratic po\-ly\-no\-mi\-als of (p,q) relatively Poisson bracket (PB). The basis of this algebra consists of elements $1,q_1,...,q_n$,\ $p_1,...,p_n$,\ $ q_i q_j, q_i p_j$,\ $p_i p_j, \quad i,j=1,...,n,\quad i\leq j$, \begin{eqnarray*} & &PB \ is \quad \{ f,g\}=\sum\frac{\partial f}{\partial p_j} \frac{\partial g}{\partial q_i}-\frac{\partial f}{\partial q_i} \frac{\partial g}{\partial p_i} \\ & &\mbox{and} \quad \{1,g \}= 0 \quad for \mbox{ all} \ g,\\ & &\{ p_i,q_j\}= \delta_{ij},\quad \{p_i q_j,q_k\}=\delta_{ik}q_j,\\ & & \{p_i q_j,p_k\}=-\delta_{jk}p_i, \quad \{p_ip_j,p_k\}=0,\\ & & \{p_i p_j,q_k \}= \delta_{ik}p_j+\delta_{jk}p_i,\quad \{ q_i q_j,q_k\}=0,\\ & & \{q_i q_j,p_k\}=-\delta_{ik}q_j-\delta_{jk}q_i \end{eqnarray*} so, we have the representation of basic elements $ f\mapsto A_f : 1\mapsto i, q_k\mapsto ix_k $, \begin{eqnarray*} && p_l\mapsto\frac{\partial}{\partial x^l}, p_i q_j\mapsto x^i\frac{\partial}{\partial x^j}+\frac{1}{2}\delta_{ij},\\ && p_k p_l\mapsto \frac{1}{i}\frac{\partial^k}{\partial x^k\partial x^l}, q_k q_l\mapsto ix^k x^l \end{eqnarray*} This gives the structure of the Poisson mani\-folds to representation of any (nilpotent) algebra or in other words to continuous wavelet trans\-form. According to this approach we can construct by using methods of geometric quantization theory many "symplectic wavelet constructions" with corresponding symplectic or Poisson structure on it. Then we may produce symplectic invariant wavelet calculations for PB or commutators which we may use in quantization procedure or in chaotic dynamics (part 8) via operator representation from section 4. \section{Quasiclassical Evolution} Let us consider classical and quantum dynamics in phase space $\Omega=R^{2m}$ with coordinates $(x,\xi)$ and generated by Hamiltonian ${\cal H}(x,\xi)\in C^\infty(\Omega;R)$. If $\Phi^{\cal H}_t:\Omega\longrightarrow\Omega$ is (classical) flow then time evolution of any bounded classical observable or symbol $b(x,\xi)\in C^\infty(\Omega,R)$ is given by $b_t(x,\xi)= b(\Phi^{\cal H}_t(x,\xi))$. Let $H=Op^W({\cal H})$ and $B=Op^W(b)$ are the self-adjoint operators or quantum observables in $L^2(R^n)$, representing the Weyl quantization of the symbols ${\cal H}, b$ [9] \begin{eqnarray*} &&(Bu)(x)=\frac{1}{(2\pi\hbar)^n}\int_{R^{2n}}b\left(\frac{x+y}{2},\xi\right) \cdot\\ &&e^{i<(x-y),\xi>/\hbar}u(y)\mathrm{d} y\mathrm{d}\xi, \end{eqnarray*} where $u\in S(R^n)$ and $B_t=e^{iHt/\hbar}Be^{-iHt/\hbar}$ be the Heisenberg observable or quantum evolution of the observable $B$ under unitary group generated by $H$. $B_t$ solves the Heisenberg equation of motion $\dot{B}_t=({i}/{\hbar})[H,B_t].$ Let $b_t(x,\xi;\hbar)$ is a symbol of $B_t$ then we have the following equation for it \begin{equation} \dot{b}_t=\{ {\cal H}, b_t\}_M, \end{equation} with the initial condition $b_0(x,\xi,\hbar)=b(x,\xi)$. Here $\{f,g\}_M(x,\xi)$ is the Moyal brackets of the observables $f,g\in C^\infty(R^{2n})$, $\{f,g\}_M(x,\xi)=f\sharp g-g\sharp f$, where $f\sharp g$ is the symbol of the operator product and is presented by the composition of the symbols $f,g$ \begin{eqnarray*} &&(f\sharp g)(x,\xi)=\frac{1}{(2\pi\hbar)^{n/2}}\int_{R^{4n}} e^{-i<r,\rho>/\hbar+i<\omega,\tau>/\hbar}\\ && \cdot f(x+\omega,\rho+\xi)\cdot g(x+r,\tau+\xi)\mathrm{d}\rho \mathrm{d}\tau \mathrm{d} r\mathrm{d}\omega. \end{eqnarray*} For our problems it is useful that $\{f,g\}_M$ admits the formal expansion in powers of $\hbar$: \begin{eqnarray*} &&\{f,g\}_M(x,\xi)\sim \{f,g\}+2^{-j}\cdot\\ &&\sum_{|\alpha+\beta|=j\geq 1}(-1)^{|\beta|}\cdot (\partial^\alpha_\xi fD^\beta_x g)\cdot(\partial^\beta_\xi gD^\alpha_x f), \end{eqnarray*} where $\alpha=(\alpha_1,\dots,\alpha_n)$ is a multi-index, $|\alpha|=\alpha_1+\dots+\alpha_n$, $D_x=-i\hbar\partial_x$. So, evolution (1) for symbol $b_t(x,\xi;\hbar)$ is \begin{eqnarray} &&\dot{b}_t=\{{\cal H},b_t\}+\frac{1}{2^j} \sum_{|\alpha|+\beta|=j\geq 1}(-1)^{|\beta|} \cdot\\ &&\hbar^j (\partial^\alpha_\xi{\cal H}D_x^\beta b_t)\cdot (\partial^\beta_\xi b_t D_x^\alpha{\cal H}).\nonumber \end{eqnarray} At $\hbar=0$ this equation transforms to classical Liouville equation \begin{equation} \dot{b}_t=\{{\cal H}, b_t\}. \end{equation} Equation (2) plays a key role in many quantum (semiclassical) problem. We note only the problem of relation between quantum and classical evolutions or how long the evolution of the quantum observables is determined by the corresponding classical one [9]. Our approach to solution of systems (2), (3) is based on our technique from [1]-[8] and very useful linear parametrization for differential operators which we present in the next section. \section{FAST WAVELET TRANSFORM FOR DIF\-FE\-RENTIAL OPERATORS} Let us consider multiresolution representation $\dots\subset V_2\subset V_1\subset V_0\subset V_{-1} \subset V_{-2}\dots$ (see our other papers from this series for details of wavelet machinery). Let T be an operator $T:L^2(R) \rightarrow L^2(R)$, with the kernel $K(x,y)$ and $P_j: L^2(R)\rightarrow V_j$ $(j\in Z)$ is projection operators on the subspace $V_j$ corresponding to j level of resolution: $(P_jf)(x)=\sum_k<f,\varphi_{j,k}>\varphi_{j,k}(x).$ Let $Q_j=P_{j-1}-P_j$ is the projection operator on the subspace $W_j$ then we have the following "microscopic or telescopic" representation of operator T which takes into account contributions from each level of resolution from different scales starting with coarsest and ending to finest scales [10]: $ T=\sum_{j\in Z}(Q_jTQ_j+Q_jTP_j+P_jTQ_j). $ We remember that this is a result of presence of affine group inside this construction. The non-standard form of operator representation [10] is a representation of an operator T as a chain of triples $T=\{A_j,B_j,\Gamma_j\}_{j\in Z}$, acting on the subspaces $V_j$ and $W_j$: $ A_j: W_j\rightarrow W_j, B_j: V_j\rightarrow W_j, \Gamma_j: W_j\rightarrow V_j, $ where operators $\{A_j,B_j,\Gamma_j\}_{j\in Z}$ are defined as $A_j=Q_jTQ_j, \quad B_j=Q_jTP_j, \quad\Gamma_j=P_jTQ_j.$ The operator $T$ admits a recursive definition via $$T_j= \left(\begin{array}{cc} A_{j+1} & B_{j+1}\\ \Gamma_{j+1} & T_{j+1} \end{array}\right),$$ where $T_j=P_jTP_j$ and $T_j$ works on $ V_j: V_j\rightarrow V_j$. It should be noted that operator $A_j$ describes interaction on the scale $j$ independently from other scales, operators $B_j,\Gamma_j$ describe interaction between the scale j and all coarser scales, the operator $T_j$ is an "averaged" version of $T_{j-1}$. The operators $A_j,B_j,\Gamma_j,T_j$ are represented by matrices $\alpha^j, \beta^j, \gamma^j, s^j$ \begin{eqnarray} \alpha^j_{k,k'}&=&\int\int K(x,y)\psi_{j,k}(x)\psi_{j,k'}(y)\mathrm{d} x\mathrm{d} y\nonumber\\ \beta^j_{k,k'}&=&\int\int K(x,y)\psi_{j,k}(x)\varphi_{j,k'}(y)\mathrm{d} x\mathrm{d} y\\ \gamma^j_{k,k'}&=&\int\int K(x,y)\varphi_{j,k}(x)\psi_{j,k'}(y)\mathrm{d} x\mathrm{d} y\nonumber\\ s^j_{k,k'}&=&\int\int K(x,y)\varphi_{j,k}(x)\varphi_{j,k'}(y)\mathrm{d} x\mathrm{d} y\nonumber \end{eqnarray} We may compute the non-standard representations of operator $\mathrm{d}/\mathrm{d} x$ in the wavelet bases by solving a small system of linear algebraical equations. So, we have for objects (4) \begin{eqnarray*} \alpha^j_{i,\ell}&=&2^{-j}\int\psi(2^{-j}x-i)\psi'(2^{-j}-\ell)2^{-j}\mathrm{d} x\\ &=&2^{-j}\alpha_{i-\ell}\\ \beta^j_{i,\ell}&=&2^{-j}\int\psi(2^{-j}x-i)\varphi'(2^{-j}x-\ell)2^{-j}\mathrm{d} x\\ &=&2^{-j}\beta_{i-\ell}\\ \gamma^j_{i,\ell}&=&2^{-j}\int\varphi(2^{-j}x-i)\psi'(2^{-j}x-\ell)2^{-j}\mathrm{d} x\\ &=&2^{-j}\gamma_{i-\ell}, \end{eqnarray*} where \begin{eqnarray*} \alpha_\ell&=&\int\psi(x-\ell)\frac{\mathrm{d}}{\mathrm{d} x}\psi(x)\mathrm{d} x\\ \beta_\ell&=&\int\psi(x-\ell)\frac{\mathrm{d}}{\mathrm{d} x}\varphi(x)\mathrm{d} x\\ \gamma_\ell&=&\int\varphi(x-\ell)\frac{\mathrm{d}}{\mathrm{d} x}\psi(x)\mathrm{d} x \end{eqnarray*} then by using refinement equations we have in terms of filters $(h_k,g_k)$: \begin{eqnarray*} \alpha_j&=&2\sum^{L-1}_{k=0}\sum^{L-1}_{k'=0}g_kg_{k'}r_{2i+k-k'},\\ \beta_j&=&2\sum^{L-1}_{k=0}\sum^{L-1}_{k'=0}g_kh_{k'}r_{2i+k-k'},\\ \gamma_i&=&2\sum^{L-1}_{k=0}\sum^{L-1}_{k'=0}h_kg_{k'}r_{2i+k-k'}, \end{eqnarray*} where $r_\ell=\int\varphi(x-\ell)\frac{\mathrm{d}}{\mathrm{d} x}\varphi(x)\mathrm{d} x, \ell\in Z.$ Therefore, the representation of $d/dx$ is completely determined by the coefficients $r_\ell$ or by representation of $d/dx$ only on the subspace $V_0$. The coefficients $r_\ell, \ell\in Z$ satisfy the following system of linear algebraical equations $$ r_\ell=2\left[ r_{2l}+\frac{1}{2} \sum^{L/2}_{k=1}a_{2k-1} (r_{2\ell-2k+1}+r_{2\ell+2k-1}) \right] $$ and $\sum_\ell\ell r_\ell=-1$, where $a_{2k-1}=$ $2\sum_{i=0}^{L-2k}h_i h_{i+2k-1}$, $k=1,\dots,L/2$ are the autocorrelation coefficients of the filter $H$. If a number of vanishing moments $M\geq 2$ then this linear system of equations has a unique solution with finite number of non-zero $r_\ell$, $r_\ell\ne 0$ for $-L+2\leq\ell\leq L-2, r_\ell=-r_{-\ell}$. For the representation of operator $d^n/dx^n$ we have the similar reduced linear system of equations. Then finally we have for action of operator $T_j(T_j:V_j\rightarrow V_j)$ on sufficiently smooth function $f$: $$ (T_j f)(x)=\sum_{k\in Z}\left(2^{-j}\sum_{\ell}r_\ell f_{j,k-\ell}\right) \varphi_{j,k}(x), $$ where $\varphi_{j,k}(x)=2^{-j/2}\varphi(2^{-j}x-k)$ is wavelet basis and $$ f_{j,k-1}=2^{-j/2}\int f(x)\varphi(2^{-j}x-k+\ell)\mathrm{d} x $$ are wavelet coefficients. So, we have simple linear para\-met\-rization of matrix representation of our differential operator in wavelet basis and of the action of this operator on arbitrary vector in our functional space. Then we may use such representation in all preceding sections. We are very grateful to M.~Cornacchia (SLAC), W.~Her\-r\-man\-nsfeldt (SLAC) Mrs. J.~Kono (LBL) and M.~Laraneta (UCLA) for their permanent encouragement.
\section{Introduction} Carraro et al (1998) developed a pure particle code, combining Barnes \& Hut (1986) octo-tree with SPH, and applying this code to the formation of a spiral galaxy like the Milky Way. The code is similar to Herniquist \& Katz (1989) TreeSPH. It uses SPH to solve the hydro-dynamical equations. In SPH a fluid is sampled using particles, there is no resolution limitation due to the absence of grids, and great flexibility thanks to the use of a time and space dependent smoothing length. Shocks are captured by adopting an artificial viscosity tensor, and the neighbors search is performed using the octo-tree. The octo-tree, combined with SPH, allows a time scaling of $N\times logN$. A good advantage of such codes is that it is easy to introduce new physics, like cooling and radiative processes, magnetic fields and so forth. Finally the kernel, which is utilised to performe hydro-dynamical quantities estimates, can be made adaptive by using anisotropic smoothing lengths. It is widely recognized that TreeSPH codes, although deficient in some aspects, can give reasonable answers in many astrophysical situations, like in simulations of fragmentation and star formation in giant molecular clouds (GMC)(), supenov\ae explosions globular clusters formation, merging of galaxies, galaxies and clusters formation and Lyman alpha forest. Galaxy formation in particular requires a huge dynamical range (Dav\'e et al 1997). In fact an ideal galaxy formation simulation would start from a volume as large as the universe to follow the initial growth of the cosmic structures, and at the same time would be able to resolve regions as small as GMC, where stars form and drive the galaxy evolution through their interaction with ISM. This ideal simulation would encompass a dynamic range of $10^{9}$ (from Gpc to parsec), $10^{6}$ time smaller than that achievable with present day codes. Big efforts have been made in the last years to enlarge as much as possible the dynamical range of numerical simulations, mainly using more and more powerful supercomputers. Scalar and vector computers indeed cannot handle efficiently a number of particles greater than half a million . Dav\'e et al (1997) for the first time developed a parallel implementation of a TreeSPH code (PTreeSPH) which can follow both collision-less and collisional matter. They report results of simulations run on a Cray T3D computer of the adiabatic collapse of an initially isothermal gas sphere (using 4096 particles), of the collapse of a Zel'dovich pancake (32768 particles) and of a cosmological simulation (32768 gas and 32768 dark particles). Their result are quit encouraging, being quite similar to those obtained with the scalar TreeSPH code (Hernquist \& Katz 1989). Porting a scalar code to a parallel machine is far from being an easy task. A massively parallel computer (like the Silicon Graphics T3E) links together hundreds or thousands of processors aiming at increasing significantly the computational power. For this reason they are very attractive, although several difficulties can arise in adapting a code to these machines. Any processor possesses its own memory, and can assess other processors memory by means of communications which are handled by an hard-ware network, and are slower than the computational speed. Great attention must be paid to avoid situations in which a small number of processors are actually working while most of them are standing idle. Usually one has to invent a proper data distribution scheme which allows to subdivide particles into processors in a way that any processor handles about the same number of particles and does not need to make heavy communications. Moreover the computational load must be shared between processors, ensuring that processors exchange informations all together, in a synchronous way, or that any processors is performing different kinds of work when it is waiting for informations coming from other processors, in an asynchronous fashion (Dav\'e et al 1997). In this paper we present a parallel implementation of the TreeSPH code dscribed in Carraro et al (1998). The numerical ingredients are the same as in the scalar version of the code. However the design of the parallel implementations required several changes to the original code. The key idea that guided us in building the parallel code was to avoid continuous communications, limiting the informations exchange at a precise moment along the code flow. This clearly reduces the communication overheard. We have also decided to tailor the code to the machine, improving its efficiency. Since we are using a T3E massively parallel computer, a natural choise was to handle communications using the SHMEN libraries, which permits asynchronous communications, and are intrinsically very fast, being produced directly by Cray for the T3E super-computer. At present the code is also portable to other machine, like SGI Origin 2000, and will be portable to any other machine with the advent of the second release of Message Passing Interface (MPI). \begin{figure*} \centerline{\epsfig{file=fig1.eps,width=12cm}} \caption{Adiabatic collapse: snapshots of the density, radial velocity, pressure and internal energy at the time of the maximum compression. The results of the test performed using $2\times 10^{4}$ particles with 8 processors are shown with dashed lines. Solid lines show the results obtained with the same number of particles, but using the scalar code, for comparison.} \end{figure*} \section{A test of the code} We consider the adiabatic collapse of an initially non-rotating isothermal gas sphere. This is a standard test for SPH codes (Hernquist \& Katz 1989. In particular it is an ideal test for a parallel code, due to the large dynamical range and high density contrast. To facilitate the comparison of our results with those by the above authors, we adopt the same initial model and the same units ($M=R=G=1$). The system consists of a $\gamma = 5/3$ gas sphere, with an initially isothermal density profile: \begin{equation} \rho(r) = \frac{M(R)}{2\pi R^{2}} \frac{1}{r} , \end{equation} \noindent where M(R) is the total mass inside the sphere of radius R. The density profile is obtained stretching an initially regular cubic grid. \begin{figure} \centerline{\epsfig{file=load.eps,width=8cm}} \caption{Overall code load-balance, averaged on 50 time-steps (solid line). Dashed line indicates ideal scalability.} \end{figure} \begin{figure} \centerline{\epsfig{file=speed.eps,width=8cm}} \caption{Overall code scalability, averaged on 50 time-steps (solid line). Dashed line indicates ideal load-balance.} \end{figure} The total number of particles used in this simulation is $2\times 10^{4}$. All the particles have the same mass. The specific internal energy is set to $u = 0.05GM/R$. For this test the viscosity parameters $\alpha$ and $\beta$ adopted are 0.5, in agreement with Dav\'e et a (1997). The gravitational softening parameter $\epsilon$ adopted for this simulation is $5 \times 10^{-3}$. The state of the system at the time of the maximum compression is shown in the various panels of Fig. 1, which displays the density, radial velocity, pressure and specific internal energy profiles. Each panel shows the variation of the physical quantity under consideration (in suitable units) as a function of the normalized radial coordinate at time equal to 0.88 . The initial low internal energy is not sufficient to support the gas cloud which starts to collapse. Approximately after one dynamical time scale a bounce occurs. The system afterwards can be described as an isothermal core plus an adiabaticlly expanding envelope pushed by the shock wave generated at the stage of maximum compression. After about three dynamical times the system reaches virial equilibrium with total energy equal to a half of the gravitational potential energy (Hernquist \& Katz 1989). The present results agree fairly well with the mean values of the Hernquist \& Katz (1989) simulations, which in turns agrees with the 1-D finite difference results. \begin{figure} \centerline{\epsfig{file=time.eps,width=12cm}} \caption{Code scalability in different code sections, averaged on 50 time-steps (solid line). Dashed line indicates ideal load-balance.} \end{figure} \begin{table*} \tabcolsep 0.1truecm \caption{The Abiabatic Collapse test. Bechmarks for a run with $2\times 10^{4}$ particles. Time refers to 50 time-stpes.} \begin{tabular}{cccccccc} \hline \multicolumn{1}{c}{$N_{cpu}$} & \multicolumn{1}{c}{Total} & \multicolumn{1}{c}{Data Up-date} & \multicolumn{1}{c}{Parallel Over-head} & \multicolumn{1}{c}{Neighbours} & \multicolumn{1}{c}{SPH} & \multicolumn{1}{c}{Gravity} & \multicolumn{1}{c}{Miscellaneous} \\ \hline & secs &secs & secs & secs & secs & secs & secs\\ \hline 1 &120 &0.47 &0.00 &40 &36 &40 &3.53\\ 2 &69 &0.22 &0.60 &23 &19 &25 &1.18\\ 4 &42 &0.27 &1.70 &14 &9.5 &15 &1.53\\ 8 &23 &0.13 &3.20 &5.5 &5.4 &5.3 &3.47\\ 16 &17.3 &0.13 &3.40 &3.4 &3.0 &3.8 &3.60\\ 32 &11.5 &0.09 &3.00 &2.6 &1.3 &3.2 &1.31\\ 64 &7.5 &0.05 &2.90 &0.33 &1.1 &1.9 &1.23\\ \hline \hline \end{tabular} \end{table*} We run the adiabatic collapse test up to the time of the maximum compression (t $\simeq 1.1$) using $2 \times 10^{4}$ particles on 1, 2, 4, 8, 16, 32 and 64 processors, and looked at the performances in the following code sections (see also Table~1): \begin{description} \item[$\bullet$] total wall-clock time; \item[$\bullet$] data up-dating data; \item[$\bullet$] parallel computation, which consists of barriers, the construction of the {\it ghost-tree} and the distribution of data between processors; \item[$\bullet$] search for neighbour particles; \item[$\bullet$] evaluation of the hydro-dynamical quantities; \item[$\bullet$] evaluation of the gravitational forces; \item[$\bullet$] miscellaneous, which encompasses I/O and kernel computation. \end{description} \noindent The results summarized in Table~1 present the total wall-clock time per timestep per processor, averaged over 50 time-steps, together with the time spent in each of the 5 subroutines (data updating, neighbour searching, SPH computation, gravitational interaction and parallel computation). The gravitation interaction takes about one-third of the total time, while the search for neighbours takes roughly comparable time. The evaluation of hydrodynamical quantities takes about one-fourth of the time, the remaining time being divided between I/O and data up-dating. The parallel over-head does not appear to be a problem, being always less than $1\%$ of the total time. This timing refers, as indicated above, to simulations stopped at roughly the time of maximum compression. A run with 8 processors up to $t \simeq 2.5$, the time at which the system is almost completely virialized, took 3800 secs. One of the most stringest requirement for a parallel code is the capability to distribute the computational work equally between all processors. This can be done defining a suitable work-load criterium, as discussed in Section~3.2. This is far from being an easy task (Dav\`e et al 1997), and in practice some processors stand idly for some time waiting that the processors with the greatest computational load accomplish their work. This is true also when an asynchronous communications scheme is adopted, as in our TreeSPH code. To evaluate the code load-balance we adopted the same strategy of Dav\`e et al (1997), measuring the fractional amount of time spent idle in a time-step while another processor performs computation: \begin{equation} L = \frac{1}{N_{procs}}\sum_{j=1}^{N_{procs}} 1 - \frac{(t_{max - t_i})}{t_{max}} . \end{equation} Here $t_{max}$ is the time spent by the slowest processor, while $t_i$ is the time taken by the $i-th$ processors to perform computation. The results are shown in Fig~2, where we plot the load-balance for simulations at increasing number of processors, from 1 to 64. The load balance maintains always above $80\%$, being close to 1 up to 8 processors. For the kind of simulation we are performing, the use of 8 processors is particulary advantageous for symmetry reasons. At increasing number of processors, a parallel code should ideally speeds up linearly In practice the increase of the processors number casuses an increase of the communications between processors, and a degradation of the code performances. To test this, we used the same simulations discussed above, ruuning the adiabatic collapse test with $2 \times 10^{4}$ particles at increasing processors number. We estimated how the code speed scales computing the wall-clock time per processor spent to execute a single time-step, averaged over 50 time-steps. In Fig.~3 we plot the speed (in $sec^{-1}$ against the number of processors. \\ The code scalabilty keeps very close to the ideal scalability up to 8 processors, where it shows a minimum. This case in fact is the most symmetricone. Then the scalability deviates significantly only when using more that 16 processors. Looking also at Fig.~4, it is easy to recognize that mainly the gravitational interaction is responsible for this deviation.\\ In the near future we are going to investigate whether the code overall scalability might improve adding new physics (like cooling and star formation) which is necessary to describe the evolution of real systems, like galaxies.
\section{Introduction} The processes that occur during the formation, evolution, and death of massive stars are the engines which drive galaxy evolution. While this statement is obvious enough, this simple fact is often lost in the quest to study the most spectacular objects or to discover the most distant galaxy. Because massive stars play such a fundamental role in galaxy evolution implies that we must have an accurate and detailed understanding of formation and evolution of massive stars and how the processes of their interaction with their surroundings changes as they evolve. Massive stars are prodigious producers of ionizing photons and mechanical energy and as such they regulate the ionization, physical, and kinematic structure of ISM. The generation of ionizing photons, mechanical energy, and the cosmic rays by massive stars regulate the ISM pressure, and by doing so, perhaps provide the mechanism by which subsequent star-formation is ultimately regulated (``feedback''). Specifically, it is this feedback from massive stars that ultimately balances star-formation against gravitational instability, tidal sheer, and dissipation that give galaxies the characteristics that they are observed to have. In all these areas, the details of stellar evolution play a critical role in understanding how stars regulate all facets of galaxy structure and evolution. Hence determining parameters like mass loss rates, time spent in various evolutionary stages, which stars go supernova, all feed back into our understanding of the structure of the ISM and hence how galaxies evolve. Of course, perhaps the best way of directly observing the role of massive stars in driving galaxy evolution is through the study of the most intense star-forming galaxies in the universe -- starbursts. The purpose of this proceeding is to review the effects of high rates of star-formation on the host galaxy's interstellar medium. Such a discussion demonstrates the central role that massive stars and the effects of stellar evolution have on the properties of galaxies. Of course, demonstrating evolutionary effects directly is difficult. We will take an indirect course. We first show that the preponderance of observational evidence is in favor of starburst galaxies driving galactic scale supernova (and stellar wind)-driven superwinds, that the strength of the wind is dependent on the star-formation rate and distribution, and then discuss the possible implications for self-regulation of star-formation, how superwinds might drive galaxy evolution, and review recent evidence that in fact, superwinds may escape the galaxy potentials in which they reside. \begin{figure}[!t] \centerline{\psfig{figure=lehnertm1.eps,width=2.5in,height=3.5in}} \vskip -0.5cm \caption{The greyscale is the H$\alpha$ image and the contours are the ROSAT PSPC image of M~82. The spatial coincidence of the H$\alpha$ and X-ray emission is quite good. The faint ridge of emission to the north is about 11.6 kpc (in projection) above the disk of M82 (see Devine and Bally 1999).} \end{figure} \section{The Basic Physics of Winds} Superwinds are thought to arise when the star-formation is intense enough to create a region of rare gas of high temperature. This hot gas has an extremely high pressure (T$\sim$10$^{7-8}$ K, n$_e$$\sim$ 0.01 -- 0.001), much higher than the ambient ISM pressure and thus is able to push the ambient ISM out preferentially in the direction of the steepest pressure gradient (i.e., the minor axis in disk galaxies). Given sufficient time and mechanical energy input, such a high pressure region will eventually ``break out'' due to the various instabilities (mainly dynamical ones) that cause the bubble walls to begin to break-up. When the bubble walls break apart, the wind begins to flow outwards eventually reaching of-order its internal sound speed as a free flowing wind (the effects of gravity can be safely ignored; see Suchkov et al 1994; Tenerio-Tagle \& Mu\~noz-Tu\~n\'on 1997). It may also shock and accelerate ambient ISM clouds to velocities of several hundred km s$^{-1}$ (Suchkov et al 1994). \begin{figure}[!t] \centerline{\psfig{figure=lehnertm2.eps,width=4.0in,height=4.0in}} \caption{In this 3 paneled figure we show the line ratios as a function of height above the disk of M82. As can be seen, the line ionization line ratios increase with projected height above the disk consistent with a greater contribution of shock-heating in the gas far above the disk.} \end{figure} The conditions for establishing such a situation are not well established theoretically and are only hinted at observationally. Obviously a necessary (but not necessarily a sufficient condition) is that the star-formation must be spatially concentrated. This allows the mechanical energy of the stellar winds and supernova remnants to be effectively and efficiently thermalized. If this mechanical energy is not thermalized efficiently, or if lots of cool material is entrained in the wind, it will radiate most of its energy away over a fairly short time scale and thus will not sustain a flow for anything like a sound crossing time over a galactic size scale (which is the minimum condition necessary for driving a galactic scale wind). Thus it is apparent that there are many factors that can keep a galactic scale wind from developing. \section{Observational Properties of Winds} Superwinds are a multi-wavelength phenomenon and the amount of observational evidence at almost all wavelengths has been growing tremendously over the past decade. Different wavelengths probe possibly different phases and phenomenology of the out-flowing wind but all show ample evidence that starburst galaxies drive superwinds (e.g., Heckman, Armus, \& Miley 1990; Lehnert \& Heckman 1996a). This evidence includes: galactic scale bi-polar spatially extended soft X-ray emission along the minor axis of starbursts disk galaxies (e.g., M82, Fig. 1; Watson, Stanger, \& Griffiths 1984; Fabbiano 1988; Bregman, Schulman, \& Tomisaka 1995; Moran \& Lehnert 1997; Ptak et al. 1997; NGC~253, Fabbiano 1988; Persic et al. 1999; NGC~1569 Heckman et al. 1995; Della Ceca et al. 1996; NGC~2146 Armus et al. 1995; Della Ceca et al. 1999; NGC~1808, Dahlem, Hartner, \& Junkes 1994; NGC~4449, Della Ceca, Griffiths, \& Heckman 1997; NGC 3628, Dahlem et al. 1996; Arp 220, Heckman et al. 1996; and for a small sample of edge-on galaxies, Dahlem, Heckman, \& Weaver 1998), extended galactic scale optical line emission with evidence for shock-heating (Fig. 2; Lehnert \& Heckman 1996a; Heckman, Armus, \& Miley 1990), emission line kinematics that often show split lines, velocity offsets relative to systemic velocity, and broad lines in the most extended emission line gas (Fig 3; Lehnert \& Heckman 1996a; Heckman, Armus, \& Miley 1990), good correlation between ionization state and line width in the extended gas (Lehnert \& Heckman 1996a), nuclear optical emission line gas and X-ray emitting plasma with extremely high pressures (several orders of magnitude higher than ambient ISM pressure in the Milky Way) and with pressure profiles that are consistent with that expected for out-flowing winds (Lehnert \& Heckman 1996a; Heckman, Armus, \& Miley 1990; Fabbiano 1988; e.g., all of the X-ray references given previously), and extended polarized radio emission (Dahlem et al. 1996). \begin{figure}[!t] \centerline{\psfig{figure=lehnertm3.eps,width=4.0in,height=4.0in}} \caption{The kinematics of the extended H$\alpha$ emission along the minor axis of M82 determined from a long slit spectrum from the KPNO 4m. The black squares denote the highest surface gas.} \end{figure} Perhaps the best and most direct probes of the wind material are X-rays. As one might recall from the previous section, the wind fluid should be hot in order to provide the high pressure necessary to drive the outflow. However, the predicted X-ray emissivity of the out-flowing material is low (Suchkov et al. 1994; 1996) and yet the observed X-ray luminosities of starburst galaxies are high (10$^{38}$ ergs s$^{-1}$ for a dwarf galaxy like NGC 1569; e.g., Heckman et al. 1995; Della Ceca et al. 1996; to 10$^{42}$ ergs s$^{-1}$ for the almost ultra-luminous IRAS galaxy NGC 3256; Moran, Lehnert, \& Helfand 1999). A number of alternatives have been proposed to enhance the X-ray luminosity produced by the wind. The wind could be centrally ``mass loaded'' whereby quantities of ambient ISM could be mixed into the wind in or near the starburst (Suchkov et al. 1996; Heckman et al. 1997). Galactic halo clouds, whether pre-existing as might be there as tidal debris of an interaction or tidal encounter or that have been carried out from the disk with the wind material itself, could be evaporated or ripped apart as they are overtaken by or interact with the wind fluid (e.g., Suchkov et al. 1994). Or the wind fluid could drive a shock into a denser volume-filling galactic halo, with the observed X-ray emission arising from the shocked-halo material rather than the wind-fluid itself. Mass loading the wind deep in the starburst nucleus would likely produce a more uniform X-ray and optical line emission in the halos of starbursts (dominated by adiabatic expansion and fluid instabilities), while interactions with halo material would be expected to produce very clumpy emission with large ranges of surface brightness and temperatures. Differentiating between the source of the halo material is less straight-forward, perhaps by using HI as a tracer of neutral material in the halo or perhaps determining the metal abundance of the high surface brightness regions of X-ray emission. \section{Conditions Necessary for Developing Outflows} The theory of superwinds suggests that to generate a wind not only requires active star-formation but in fact that the star-formation must be intense. Intense in this context means that the volume density of energy input must be high enough so that the mechanical energy deposited by stars into the ISM is effectively thermalized before it has a chance to radiate away a significant fraction of its mechanical energy. Lehnert \& Heckman (1996) and Heckman, Armus, \& Miley (1990) in studies of galaxies selected from the IRAS survey (i.e., infrared bright) found that it is not only the star-formation rate but other factors like the ``warmth'' of the IR emission and the ratio of IR to optical luminosity also influenced the observational strength of the wind. Both of these additional factors are related to how enshrouded the starburst region is and the UV heating rate of the dust (which are proportional in some sense to the volume density of the energy input). These results substantiate one of the basic tenets of the superwinds hypothesis. Specifically what these studies found was that to drive an outflow, a galaxy should have large IR luminosities (L$_{IR}$ $>$ 10$^{44}$ erg s$^{-1}$), large IR excesses (L$_{IR}$/L$_{OPT}$ $>$ 2), and warm far-IR colors ($S_{60\mu m}/S_{100\mu m} \geq 0.5$). Taking these limits, the IR luminosity function, outflow rates $\propto$ L$_{IR}$ and using constants of proportionality determined in well-studied examples like M82 and NGC253, the local space density of galaxies, and a value of the Hubble time, we derive that superwinds have carried out: M$_{eject}$ $\approx$ 5 $\times$ 10$^8$ M$_{\sun}$ in metals, and E$_{KE \ and \ Thermal}$ $\approx$ 10$^{59}$ ergs per average (Schecter L$^{\star}$) galaxy over the history of the universe. Interestingly, these are approximately the mass of metals and the binding energy of an average (L$^{\star}$) galaxy. And this estimate is conservative in that it assumes no evolution in the star-formation rate with epoch. Reasonable assumptions about the increasing interaction rate and starburst number density with epoch would only increase our estimates and demonstrates the potentially substantial role outflows have played in galaxy formation and evolution. \section{Questions Surrounding and Implications of Superwinds} There are two central questions that studying superwind engenders. Is there some critical star-formation intensity at which the mechanical energy output from the massive stars halts further star-formation? Does the wind ultimately escape the potential of the host galaxy? Interestingly, it appears that the answer to the first question may be yes. For example, in a study of IR selected starbursts, Lehnert \& Heckman (1996b) found that these galaxies have a ``large scale'' IR surface brightness that appears to have a limit of $\approx$10$^{11}$ L$_{\sun}$ kpc$^{-2}$. In a similar study but now including starbursts with a range of selection methods and a wide range of redshifts, Meurer et al. (1997) found a similar limit of L$_{BOL}$$\approx$ 2 $\times$ 10$^{11}$ L$_{\sun}$ kpc$^{-2}$ (see Weedman et al. 1998 who suggest that this limit may increase with increasing redshift -- about a factor of 4 at redshifts $\approx$2--3). This corresponds to a maximum star-formation rate of about 45 M$_{\sun}$ yr$^{-1}$ kpc$^{-2}$ for a ``normal'' initial mass function. This result should not be over-interpreted. It is not to mean that in all cases, star-formation is stopped at this limit. The cores of some HII regions and the super-star clusters violate this ``limit''. This is only to suggest that perhaps the feedback from massive stars might provide a global integrated limit to the star-formation rate. On smaller scales, different physical processes undoubtedly dominate compared to those on larger scales which can lead to very high star-formation rates per unit area or volume over limited scales. In an analysis of this problem, Lehnert \& Heckman (1996b) argued that this limit could plausibly be due to the out-flowing wind providing enough pressure to overcome the hydrostatic pressure and thus halt further star-formation. To show this is plausible, they used M82 as a test case and showed that star-formation in M82, which is at or near this limit in the star-formation rate per unit area, does provide enough pressure in the wind to balance the hydrostatic pressure of the disk of M82. However, such an hypothesis awaits further testing (see also the analysis of Meurer et al. 1997). With the rapid decline in the surface brightnesses of spatially extended X-ray, optical, and radio emission, answering the question of whether or not the wind fluid ultimately escapes the galactic potential is currently a difficult question to answer definitively. Certainly, the model prediction is that it should, but whether or not it actually does depends on where and how the wind is mass-loaded and how much cool material ultimately mixes with the wind fluid. However, some hints at the answer are starting to emerge. For example, Norman et al. (1996) in a study of absorption lines in QSOs projected near the line of sight to NGC 520 and NGC 253 found that it is plausible that the absorption lines they detected along the line of sight to NGC 520 could be associated with an outflow thus suggesting that the wind could reach large distances from the galaxy. However, as they pointed out, this is not a unique interpretation of their results. Another possible way of addressing this question has recently presented itself. Devine and Bally (1999) have recently discovered a ridge of correlated X-ray and H$\alpha$ line emission 11.6 Kpc (in projection) above the disk of M82 (Fig. 1). In an analysis of this ridge of emission, Lehnert, Heckman, \& Weaver (1999, submitted to ApJ) have argued that this ridge represents the interaction of the superwind with a dense cloud in the halo of M82 that may be part of the tidal debris from the interaction between M82/NGC3077/M81 (Yun et al. 1993; 1994). The analysis reveals that in order to explain the X-ray properties of the ridge of X-ray/H$\alpha$ emission, the out-flowing wind must be hitting the cloud at about 800 km s$^{-1}$ -- well in excess of the escape speed at that distance above the plane for M82. There is little doubt that the wind in M82 is escaping its galactic potential. \section{Conclusions} We have demonstrated that the preponderance of evidence is strongly in favor of starburst driving galactic scale superwinds. In spite of the fact that the existence of superwind is on a firm observational basis, there are still many unanswered questions. We have suggested that winds are likely to have a huge impact during galaxy formation and subsequent evolution but we do not understand the details of that statement. We know that the feedback from massive stars is likely to be important, but how important? What is the nature of the X-ray emission seen in starburst galaxies? The wind in M82 seems to be able to escape the galactic potential, but is the escape of the wind plasma a general feature of galaxies with superwinds? Was it easier or more difficult in the past for galaxies to drive winds and for these winds to escape the galactic potential? How will the details of stellar evolution and the mechanical energy input change our views of winds and their influence on galaxy evolution? Some of these questions can only be answered definitively when we have a complete understanding of the physics and evolution massive stars including the role of Wolf-Rayet stars in exciting and disturbing the ISM. \acknowledgments I would like to express my sincerest thanks to Karel van der Hucht for his immense patience and understanding in waiting for my contribution.
\section{Introduction} It has been established that at low energies D-branes are described effectively by the Born-Infeld action, whose lowest order with respect to the number of derivatives is the Super-Yang-Mills theory (SYM) \cite{Witten}. When M-theory is compactified in the infinite momentum frame, only the degrees of freedom of D-particles remain. It was therefore conjectured that the system is fundamentally described by $d=1$ SYM, i.e., the Matrix model \cite{BFSS,Susskind}. The conjecture of the Matrix model has been intensively investigated in many systems and compared to supergravity. Especially, D-particle scatterings with multi-bodies have been checked up to two loops \cite{DKPS,BBPT,OY}, though disagreement for higher loops in a more complicated system is suspected \cite{DEG}. Therefore, we have to show the full agreement between the Matrix model and supergravity to confirm the validity of the Matrix model as a non-perturbative definition of M-theory. A possible way to show the full agreement is to rely on symmetries. For example, the supersymmetries put restrictions on the form of the Matrix model effective action in the first few orders of the derivative expansion \cite{SSP,SC,O}. Conformal symmetry imposes another restriction on the effective action of the SYM, and in particular, the Matrix model. In fact, it is shown in \cite{Malda} that the Born-Infeld action with the background of the near-horizon geometry of the D3-brane solution (that is, the Anti-de-Sitter or AdS space) can be determined exactly by the isometry of the AdS space. The special conformal transformation (SCT) of the isometry of the AdS space differs from the canonical one in SYM by an extra term which vanishes on the boundary. However, interestingly, this extra term can be derived from SYM as a quantum correction \cite{JKY3}. In this way, the problem of showing the full agreement between the Born-Infeld action with the AdS background and the effective action of SYM is reduced to showing that the quantum modified symmetry of SYM reproduces exactly the isometry of the AdS space. Although the maximally SYM in $d\ne 4$ is not a conformal field theory and the near-horizon geometry of the D$p$-brane with $p\ne 3$ is not of the AdS type, one can generalize the above arguments to D$p$-branes by varying also the coupling constant under the dilatation and the SCT in both the SYM and the D$p$-brane geometry \cite{JY}. This time, though the isometry of the near-horizon geometry determines the Born-Infeld action, the SCT derived from the SYM appears to differ from that of the isometry by a numerical factor. The SCT Ward-Takahashi identity still holds only when one keeps the terms proportional to the derivative of the coupling constant in the effective action, because they are also relevant after the transformation \cite{JKY0,HM}. It is also shown that the two SCTs (i.e., the isometry and the quantum modified one in SYM) are related by a field redefinition. Though consistent, the notorious numerical factor prevents us from determining the Born-Infeld action directly. Hence, it would be difficult to show the full agreement between the Matrix model and the Born-Infeld theory from the symmetry. Besides, it is unclear why the field redefinition is necessary in spite of the fact that in other works \cite{DKPS,BBPT,OY} studying the relationship between the Matrix model and the supergravity they always agree without any field redefinitions. In this paper, we present a new gauge in SYM, which is a natural extension of the background gauge adopted in all the previous works \cite{JY,JKY0,HM}. This new gauge has a marvelous property that the SCT as the isometry of the near-horizon geometry of the D$p$-brane solution is correctly reproduced without field redefinitions as the quantum modified SCT of SYM, unlike in the case of the conventional background gauge. Then, the result in our new gauge raises a question: what is the meaning of the gauge choice in the Matrix model? Therefore, we give a general argument to identify the change of gauge-fixing in the Matrix model as the redefinition of the D-brane coordinates. As a concrete example of this argument, we carry out the calculation in the $R_\xi$ gauge. We find among other things that the agreement between the Matrix model and the supergravity without any field redefinitions \cite{DKPS,BBPT,OY} is merely a special nature of the background gauge. The organization of the rest of this paper is as follows. In section 2, we summarize the quantum conformal symmetry in SYM. We derive the quantum modified SCT and explain the disagreement of the numerical factor with that in the isometry. In section 3, we present our new gauge and analyze the quantum SCT in this gauge. In section 4, we give a general argument on the relation of changing the gauge in SYM to the field redefinition in the effective action, and in section 5 we analyze the $R_\xi$ gauge as an example of the general formalism given in section 4. In the final section, we summarize the paper and discuss further directions. In the appendices, we present some technical details used in the text. \section{Quantum conformal symmetry in SYM} First, let us summarize the quantum conformal symmetry in SYM \cite{JY,JKY3,JKY0,HM} in detail, because the methods needed afterwards are essentially the same. It is well known that the action of $d=4$ ${\cal N}=4$ SYM has conformal symmetry. This is also the case for the SYM with sixteen supersymmetries in other dimensions if we assign the coupling constant a conformal dimension and vary it under the transformation. We will here consider SYM in the Euclidean formulation: \begin{equation} S_{\rm SYM}=\int\!{\rm d}^{p+1}\!x\,\frac{1}{{g_{\rm YM}^2}} \mathop{\rm tr}\left(\frac{1}{4}F_{\mu\nu}^2 +{1 \over 2}(D_\mu X_m)^2-\frac{1}{4}[X_m,X_n]^2+ \mbox{fermionic part} \right). \label{SYM} \end{equation} The action (\ref{SYM}) is invariant under both the dilatation and SCT. In particular, the transformation law of the bosonic variables under SCT reads \begin{eqnarray} &&\delta_{\rm SCT}^{\rm C} x_\mu=2\epsilon\!\cdot\! xx_\mu-\epsilon_\mu x^2,\quad \delta_{\rm SCT}^{\rm C} A_\mu=-2\epsilon\!\cdot\! xA_\mu -2(x\!\cdot\! A\epsilon_\mu-\epsilon\!\cdot\! Ax_\mu),{\nonumber}\\ &&\delta_{\rm SCT}^{\rm C} X_m=-2\epsilon\!\cdot\! xX_m,\quad \delta_{\rm SCT}^{\rm C}{g_{\rm YM}^2}=2(p-3)\epsilon\!\cdot\! x{g_{\rm YM}^2},\label{SCT} \end{eqnarray} where the superscript C in $\delta_{\rm SCT}^{\rm C}$ is for distinguishing the present classical transformation from the quantum one to be given later. In order to quantize the system, we have to add the gauge-fixing and the corresponding ghost terms to the original action (\ref{SYM}): \begin{eqnarray} S_{\rm gf+gh}=\int\!{\rm d}^{p+1}\!x\,i\delta_{\rm BRST}\mathop{\rm tr}(\bar{C}G). \label{gf+gh} \end{eqnarray} In the original work of \cite{JKY0}, they adopted the famous background gauge \begin{equation} G=-{\partial}_\mu A_\mu+i[B_m,Y_m]+{1 \over 2}{g_{\rm YM}^2} b, \label{BGgauge} \end{equation} where $Y_m$ is the fluctuation of the scalars $X_m$ from the diagonal background $B_m$; $X_m=B_m+Y_m$, and $b$ is the auxiliary field for the off-shell closure of the BRST algebra. The BRST transformation of the fields are \begin{eqnarray} &&\delta_{\rm BRST} A_\mu= D_\mu C\equiv {\partial}_\mu C -i[A_\mu,C],\quad \delta_{\rm BRST} X_m=-i[X_m,C],{\nonumber}\\ &&\delta_{\rm BRST} C=i C^2,\quad \delta_{\rm BRST}\bar{C}=i b,\quad \delta_{\rm BRST} b=0. \label{dBRST} \end{eqnarray} We assign the SCT of the unphysical fields so that $\delta_{\rm SCT}^{\rm C}$ and the BRST transformation $\delta_{\rm BRST}$ are commutative, $[\delta_{\rm SCT}^{\rm C},\delta_{\rm BRST}]=0$: \begin{eqnarray} \delta_{\rm SCT}^{\rm C} C=0,\quad \delta_{\rm SCT}^{\rm C}\bar{C}=-2(p-1)\epsilon\!\cdot\! x\,\bar{C},\quad \delta_{\rm SCT}^{\rm C} b=-2(p-1)\epsilon\!\cdot\! x\,b. \label{unphysSCT} \end{eqnarray} While $S_{\rm gf+gh}$ (\ref{gf+gh}) with $G$ of (\ref{BGgauge}) is dilatation invariant, it is not invariant under SCT (\ref{SCT}): \begin{eqnarray} \delta_{\rm SCT}^{\rm C} S_{\rm gf+gh}=i\delta_{\rm BRST}\lambda[A,\bar{C}], \label{dltS} \end{eqnarray} where $\lambda[A,\bar{C}]$ is given by \begin{eqnarray} \lambda=\frac{p-1}{2}\!\cdot\!(-4)\int\!{\rm d}^{p+1}\!x \mathop{\rm tr}\left(\bar{C}(x)A(x)\!\cdot\!\epsilon\right). \label{lambda} \end{eqnarray} In general, if a symmetry of the classical action is violated by the gauge-fixing and the ghost terms $S_{\rm gf+gh}$ and the violation is given as the BRST-exact form $i\delta_{\rm BRST} \lambda$, the symmetry can be restored by adding to the original transformation for a generic field $\phi$ the BRST transformation $-i\lambda\delta_{\rm BRST}\phi$ with this (field dependent) transformation parameter $\lambda$. In fact, the change of the path-integral measure under the added BRST transformation just cancels the violation $i\delta_{\rm BRST} \lambda$. Therefore, in the present case, the effective action $\Gamma[B,{g_{\rm YM}^2}]$ of the system satisfies the following SCT Ward-Takahashi identity, \begin{equation} \int\!{\rm d}^{p+1}\!x\left(\delta_{\rm SCT}^{\rm C} {g_{\rm YM}^2}(x)\frac{\delta}{\delta {g_{\rm YM}^2}(x)} +\left(\delta_{\rm SCT}^{\rm C}+\delta_{\rm SCT}^{\rm Q}\right)\!B_{m,i}(x)\,\frac{\delta}{\delta B_{m,i}(x)} \right)\Gamma[B,{g_{\rm YM}^2}]=0, \label{WI} \end{equation} where the extra term $\delta_{\rm SCT}^{\rm Q} B_{m,i}$ for $B_m=\mathop{\rm diag}\left(B_{m,i}\right)$ is \begin{eqnarray} \delta_{\rm SCT}^{\rm Q} B_{m,i}(x)=2(p-1)\Bigl\langle[C(x),X_m(x)]_{ii} \int\!{\rm d}^{p+1}y\,\mathop{\rm tr}(\bar{C}(y)A(y)\!\cdot\!\epsilon)\Bigr\rangle. \label{extraterm} \end{eqnarray} Note that the total SCT for $B_m$ is now given as a sum of the classical part $\delta_{\rm SCT}^{\rm C} B_{m,i}=-2\epsilon\cdot xB_{m,i}$ and the quantum correction $\delta_{\rm SCT}^{\rm Q} B_{m,i}$. Now let us explicitly calculate $\delta_{\rm SCT}^{\rm Q} B_{m,i}$ (\ref{extraterm}). At the 1-loop order it is given by \begin{eqnarray} \delta_{\rm SCT}^{\rm Q} B_{m,i}(x)=2(p-1)\int\!{\rm d}^{p+1}\!y\left( \langle C_{ij}(x)\,\bar{C}_{ji}(y)\rangle \VEV{Y_{m,ji}(x)A_{\mu,ij}(y)\epsilon_\mu} -(i\leftrightarrow j)\right), \label{extraterm_oneloop} \end{eqnarray} where the free propagators are \begin{eqnarray} \langle C_{ij}(x)\bar{C}_{ji}(y)\rangle\!\!\!&=&\!\!\! i\bra{x}\Delta_{ij}\ket{y},{\nonumber}\\ \VEV{Y_{m,ji}(x)A_{\mu,ij}(y)}\!\!\!&=&\!\!\! -2i{g_{\rm YM}^2}\bra{x}\Delta_{ij}({\partial}_\nu B_{m,ij})\Delta_{ij} \left({\cal M}^{-1}\right)_{\nu\mu}\ket{y},\label{freeprop} \end{eqnarray} with $\Delta_{ij}\equiv\left(-{\partial}^2+B_{ij}^2\right)^{-1}$, $B_{m,ij}\equiv B_{m,i}-B_{m,j}$ and ${\cal M}_{\mu\nu}\equiv\delta_{\mu\nu} -4({\partial}_\mu B_{\ell,ij})\Delta_{ij}({\partial}_\nu B_{\ell,ij})\Delta_{ij}$. Keeping only the lowest order terms in the derivatives, eq.\ (\ref{extraterm_oneloop}) is reduced to \begin{equation} \delta_{\rm SCT}^{\rm Q} B_{m,i}(x)=\sum_{j} 8(p-1){g_{\rm YM}^2}\epsilon\!\cdot\!{\partial} B_{m,ij}\Delta_{ij}^3 =\sum_{j}\frac{4(p-1)\Gamma\Bigl((5-p)/2\Bigr){g_{\rm YM}^2}} {(4\pi)^{(p+1)/2}B_{ij}^{5-p}}\epsilon\!\cdot\!{\partial} B_{m,ij}. \label{Dlt_B_final} \end{equation} Restricting ourselves to the source-probe configuration with $N$ D$p$-branes as the source at the origin and the probe at $B_m$; $B_m=\mathop{\rm diag}(0,\cdots,0,B_m)$, we obtain the final form of the quantum modified SCT $\delta_{\rm SCT}\equiv\delta_{\rm SCT}^{\rm C}+\delta_{\rm SCT}^{\rm Q}$ for $x_\mu$, ${g_{\rm YM}^2}$ and $U_m\equiv 2\pi B_m$: \begin{eqnarray} \delta_{\rm SCT} x_\mu\!\!\!&=&\!\!\! 2\epsilon\!\cdot\! x\,x_\mu-\epsilon_\mu x^2,{\nonumber}\\ \delta_{\rm SCT} {g_{\rm YM}^2}\!\!\!&=&\!\!\! 2(p-3)\epsilon\!\cdot\! x\,{g_{\rm YM}^2},{\nonumber}\\ \delta_{\rm SCT} U_m\!\!\!&=&\!\!\!-2\epsilon\!\cdot\! x\,U_m -\frac{p-1}{2}\frac{kR_p^4}{U^2}\epsilon\!\cdot\!{\partial}\,U_m, \label{finalsct} \end{eqnarray} with \begin{eqnarray} k\equiv\frac{2}{5-p},\quad R_p^2\equiv\sqrt{d_p{g_{\rm YM}^2} NU^{p-3}},\quad d_p\equiv 2^{7-2p}\pi^{(9-3p)/2}\Gamma\Bigl(\frac{7-p}{2}\Bigr). \end{eqnarray} The quantum part $\delta_{\rm SCT}^{\rm Q} U_m$ of SCT has an extra numerical factor $(p-1)/2$ compared to the isometry of the near-horizon geometry of the D$p$-brane solution \cite{JKY0}. Since it is the isometry of the near-horizon geometry that determines the Born-Infeld action with the background, one may wonder if the SCT derived in SYM is consistent with the Ward-Takahashi identity. However, it was pointed out in \cite{JKY0} that, since the derivative of the coupling constant $\eta_\mu\equiv{\partial}_\mu{g_{\rm YM}^2}/{g_{\rm YM}^2}$ transforms under SCT as \begin{eqnarray} \delta_{\rm SCT}\eta_\mu=2(p-3)\epsilon_\mu+{\cal O}(\eta), \end{eqnarray} we have to keep terms linear in $\eta_\mu$ in the calculation of the 1-loop effective action to confirm the validity of the Ward-Takahashi identity.\footnote{ We consider the lowest non-trivial order in $\eta_\mu$ and hence put $\eta_\mu=0$ after the SCT.} Let us check it for the D-particle case. For this purpose, we need the quadratic parts of the action: \begin{eqnarray} &&{\cal L}_{YY}={1\over 2{g_{\rm YM}^2}} Y_{m,ij}(-{\partial}^2+\eta{\partial}+B_{ij}^2)Y_{m,ji},\quad {\cal L}_{AA}={1\over 2{g_{\rm YM}^2}} A_{ij}(-{\partial}^2+\eta{\partial}+B_{ij}^2)A_{ji},{\nonumber}\\ &&{\cal L}_{YA}={2i\over {g_{\rm YM}^2}}V_{m,ij}Y_{m,ij}A_{ji},\quad {\cal L}_{\bar{C}C}= -i\bar{C}_{ij}(-{\partial}^2+B_{ij}^2)C_{ji},\label{quadratic}\\ &&{\cal L}_{\theta\theta}={1 \over 2}\theta_{\alpha,ij} (-\delta_{\alpha\beta}{\partial}+\gamma^m_{\alpha\beta}B_{m,ij}) \theta_{\beta,ji},{\nonumber} \end{eqnarray} with $V_{m,ij}\equiv\dot B_{m,ij}-{1 \over 2}\eta B_{m,ij}$. We can read off the 1-loop effective action for the source-probe situation from (\ref{quadratic}) as \begin{eqnarray} \Gamma_{1\hbox{\scriptsize -loop}}\!\!\!&=&\!\!\! N\mathop{\rm Tr}\Bigl\{10\ln\Bigl(-{\partial}^2+\eta{\partial}+B^2\Bigr) +\ln\Bigl(1-4V_m\Delta^{\eta} V_m\Delta^{\eta}\Bigr){\nonumber}\\ &&\quad-2\ln\Bigl(-{\partial}^2+B^2\Bigr) -4\sum_\pm\ln\left(-{\partial}^2+B^2\pm\dot{B}\right)\Bigr\}, \label{one-loop} \end{eqnarray} where $\mathop{\rm Tr}$ denotes the trace over the functional space of $\tau$, and $\Delta^{\eta}$ is defined by $\Delta^{\eta}\equiv(-{\partial}^2+\eta{\partial}+B^2)^{-1} =\Delta-\Delta\eta{\partial}\Delta+{\cal O}(\eta^2)$ with $\Delta\equiv(-{\partial}^2+B^2)^{-1}$. In (\ref{one-loop}), the first term is the contribution of ${\cal L}_{YY}$ and ${\cal L}_{AA}$, the second term is due to the mixing between $Y_m$ and $A$, the third term is the ghost loop and the last term is from ${\cal L}_{\theta\theta}$. Keeping only terms independent of and linear in $\eta$, we have \begin{eqnarray} \Gamma_{1\hbox{\scriptsize -loop}}\!\!\!&=&\!\!\! N\mathop{\rm Tr}\Bigl\{8\ln\Bigl(-{\partial}^2+B^2\Bigr) +\ln\Bigl(1-4\dot B_m\Delta\dot B_m\Delta\Bigr) -4\sum_\pm\ln\Bigl(-{\partial}^2+B^2\pm\dot{B}\Bigr){\nonumber}\\ &&+10\eta{\partial}\Delta+4\eta\Bigl(B_m\Delta\dot B_m\Delta +2{\partial}\Delta\dot B_m\Delta\dot B_m\Delta\Bigr) \Bigl(1-4\dot B_\ell\Delta\dot B_\ell\Delta\Bigr)^{-1}\Big\}. \end{eqnarray} Note that the SCT of the terms of the form $\int{\rm d}\tau\eta(\tau)\times\mbox{(total derivative terms)}$ in the effective action vanish if we put $\eta=0$ after the transformation. Then, since we have $\bra{\tau}{\partial}{\cal O}\ket{\tau}=(1/2){\partial}_\tau\!\bra{\tau}{\cal O}\ket{\tau}$ for ${\cal O}=\Delta$ and $\Delta\dot B_m\Delta\dot B_m\Delta (1-4\dot B_\ell\Delta\dot B_\ell\Delta)^{-1}$ due to $\bra{\tau_1}\Delta\ket{\tau_2}=\bra{\tau_2}\Delta\ket{\tau_1}$, the terms proportional to $\eta$ and relevant for SCT are \begin{eqnarray} \mathop{\rm Tr}\Bigl\{4\Bigl(\eta B_m\Delta\dot B_m\Delta\Bigr) \Bigl(1-4\dot B_\ell\Delta\dot B_\ell\Delta\Bigr)^{-1}\Big\} =\mathop{\rm Tr}\Bigl\{4\eta B_m\Delta\dot B_m\Delta+16\eta B_m\Delta\dot B_m\Delta\dot B_\ell\Delta\dot B_\ell\Delta\Bigr\}, \label{etalinear} \end{eqnarray} where we have kept only the terms with the number of derivatives less than or equal to four. Note that we cannot adopt the eikonal approximation and drop the acceleration terms here \cite{HM}, because by integration by parts, they can be converted to terms without the accelerations. A method to evaluate it was given in our previous work \cite{HM}: we took polynomial forms for the background $B_m(\tau)$, calculated the 1-loop effective action, and identified the result as a functional of $B_m(\tau)$. An equivalent, but more refined method was presented in \cite{O}: all the terms are rearranged into the forms of $f(\tau){\partial}^m\Delta^n$, which are calculated using the proper-time representation. Here we adopt the more convenient method of \cite{O}. Using the formulas presented in the appendix, the terms linear in $\eta$ in $\Gamma_{1\hbox{\scriptsize -loop}}$ are calculated to give \begin{eqnarray} N\int\!{\rm d}\tau\eta\biggl(\frac{1}{4}\frac{\dddot{B}\!\cdot\! B}{B^5} -\frac{5}{2}\frac{\ddot B\!\cdot\! B\dot B\!\cdot\! B}{B^7} +\frac{5}{4}\frac{\dot B^2\dot B\!\cdot\! B}{B^7} +\frac{35}{8}\frac{(\dot B\!\cdot\! B)^3}{B^9} +{\rm total\,\,derivative\,\,terms}\biggr). \end{eqnarray} After making further the identification of the total derivative terms, we find the final expression of $\Gamma_{1\hbox{\scriptsize -loop}}$: \begin{eqnarray} \Gamma_{1\hbox{\scriptsize -loop}} \!\!\!&=&\!\!\! N\int\!{\rm d}\tau\biggl\{-\frac{15}{16}\frac{(\dot{B}^2)^2}{B^7} +\eta\biggl(\frac{15}{8}\frac{\dot{B}^2\dot{B}\!\cdot\! B}{B^7} +{\rm total\,\,derivative\,\,terms}\biggr)\biggr\}. \label{Goneloop} \end{eqnarray} The problem of the extra factor $(p-1)/2$ in $\delta_{\rm SCT}^{\rm Q} U_m$ (\ref{finalsct}) is now resolved by taking into account the SCT of the term $\eta(15/8)\dot B^2\dot B\!\cdot\! B/B^7$ in (\ref{Goneloop}). Moreover, this $\eta$-dependent term in (\ref{Goneloop}) can be eliminated by making the field redefinition $B_m\to\widetilde{B}_m$ with \begin{equation} \widetilde B_m=B_m-{g_{\rm YM}^2} N\frac{3}{4}\frac{\eta\dot B_m}{B^5}, \label{tildeB} \end{equation} in $\Gamma_{\rm tree}=\int{\rm d}\tau\dot B_m^2/2{g_{\rm YM}^2}$. Then, the SCT for the new variable $\widetilde B_m$ is that of the isometry without the extra factor $(p-1)/2$. \section{A new gauge} Instead of the usual background gauge (\ref{BGgauge}), here we propose a bit different gauge function useful for discussing the conformal symmetry in the SYM. In fact, we shall find that the notorious numerical factor $(p-1)/2$ does not appear this time. The gauge function of our new gauge is \begin{eqnarray} G=-{\partial}_\mu A_\mu+\eta_\mu A_\mu+i[B_m,Y_m]+{1 \over 2}{g_{\rm YM}^2} b, \label{NEWgauge} \end{eqnarray} which has the additional term $\eta_\mu A_\mu$ compared to the old one (\ref{BGgauge}). As in the previous case, the SCT symmetry broken by $S_{\rm gf+gh}$ can be restored by adding to the classical SCT the BRST transformation with a field dependent parameter $\lambda$. In our new gauge, $\lambda[\bar{C},A]$ is \begin{eqnarray} \lambda=-4\int\!{\rm d}^{p+1}\!x\mathop{\rm tr}\Bigl(\bar{C}(x)A(x)\!\cdot\!\epsilon\Bigr). \label{lambdanew} \end{eqnarray} Note that in the usual background gauge (\ref{BGgauge}) the factor $(p-1)/2$ in $\delta_{\rm SCT}^{\rm Q} U_m$ (\ref{finalsct}) originates in (\ref{lambda}). However, it is missing in (\ref{lambdanew}). This implies that we can derive the SCT of the isometry with the correct factor in our new gauge, namely, $\delta_{\rm SCT}^{\rm Q} U_m$ in the new gauge is given by \begin{equation} \delta_{\rm SCT} U_m=-2\epsilon\!\cdot\! x\,U_m -\frac{kR_p}{U^2}\epsilon\!\cdot\!{\partial}\,U_m, \end{equation} instead of that in (\ref{finalsct}). Since now the isometry that determines the Born-Infeld action has been reproduced with the correct coefficient, we can expect that the $\eta$-dependent terms such as $\eta(15/8)\dot B^2\dot B\!\cdot\! B/B^7$ in (\ref{Goneloop}) are missing from the one-loop effective action for the D-particle. Let us explicitly check it in the rest of this section. The quadratic parts of the D-particle action in the new gauge are \begin{eqnarray} &&{\cal L}_{YY}={1\over 2{g_{\rm YM}^2}}Y_{m,ij}(-{\partial}^2+\eta{\partial}+B_{ij}^2)Y_{m,ji}, \quad {\cal L}_{AA}={1\over 2{g_{\rm YM}^2}}A_{ij}(-{\partial}^2-\eta{\partial}+B_{ij}^2)A_{ji},{\nonumber}\\ &&{\cal L}_{YA}={2i\over {g_{\rm YM}^2}}\dot B_{m,ij}Y_{m,ij}A_{ji},\quad {\cal L}_{\bar{C}C}=-i\bar{C}_{ij}(-{\partial}^2+\eta{\partial}+B_{ij}^2)C_{ji}, \label{newgaugequadratic} \end{eqnarray} and ${\cal L}_{\theta\theta}$ is the same as (\ref{quadratic}) in the old gauge. The one-loop effective action for the source-probe configuration is given by \begin{eqnarray} &&\Gamma_{1\hbox{\scriptsize -loop}}= N\mathop{\rm Tr}\Bigl\{9\ln\left(-{\partial}^2+\eta{\partial}+B^2\right) +\ln\left(-{\partial}^2-\eta{\partial}+B^2\right){\nonumber}\\ &&\qquad\qquad +\ln\Bigl(1-4\dot B_m\bigl(-{\partial}^2+\eta{\partial}+B^2\bigr)^{-1} \dot B_m\bigl(-{\partial}^2-\eta{\partial}+B^2\bigr)^{-1}\Bigr){\nonumber}\\ &&\qquad\qquad -2\ln\left(-{\partial}^2+\eta{\partial}+B^2\right) -4\sum_{\pm}\ln\Bigl(-{\partial}^2+B^2\pm\dot{B}\Bigr)\Bigr\}, \end{eqnarray} where the origin of the respective terms are the same as before (\ref{one-loop}). The $\eta$-independent term in $\Gamma_{1\hbox{\scriptsize -loop}}$ is the same as in (\ref{Goneloop}), while the term linear in $\eta$ is seen to be expressed as \begin{equation} \mathop{\rm Tr}\Bigl\{\eta{\partial}{\cal O}\Bigr\} =\int\!{\rm d}\tau\,\eta(\tau){1 \over 2}{\partial}_\tau\!\bra{\tau}{\cal O}\ket{\tau}, \label{etaterm} \end{equation} in terms of a symmetric ${\cal O}$ satisfying $\bra{\tau_1}{\cal O}\ket{\tau_2}=\bra{\tau_2}{\cal O}\ket{\tau_1}$. Eq.\ (\ref{etaterm}) is SCT invariant by putting $\eta=0$ after the transformation. Let us summerize our findings in this section. In the usually adopted background gauge (\ref{BGgauge}), the quantum SCT has an extra numerical factor $(p-1)/2$. However, if we adopt the new gauge (\ref{NEWgauge}), this extra factor disappears and we can derive directly the conformal symmetry that determines the full Born-Infeld action. Accordingly, the $\eta$-dependent terms in $\Gamma_{1\hbox{\scriptsize -loop}}$ is also missing up to total derivative terms in our new gauge. The calculation in this section shows that the form of quantum SCT depends crucially on the choice of gauge. Therefore one may question what the role of the gauge function is. We can find a clue in \cite{JKY0}, where it is shown that the SCT with the extra factor $(p-1)/2$ can be related to the SCT without it by the field redefinition (\ref{tildeB}). Therefore it seems that the role of the gauge function is to choose the definition of the fields in the Born-Infeld action side. In the following sections, we shall study the relationship between the gauge choice and the redefinition of the fields. \section{Gauge shifts and field redefinitions} Here we give some general arguments on the relations between the change of the gauge function and the field redefinition. The following is essentially the reproduction of the old arguments about the independence of the physical S-matrices in the Yang-Mills theory on the choice of gauge \cite{Lee}. Let us consider the partition function, \begin{eqnarray} Z_G[J]=\int\!{\cal D}\phi\,e^{-S+J\cdot\phi}, \end{eqnarray} where $\phi$ denotes collectively all the fields in the system and $J$ is the corresponding source. The action $S$ consists of the gauge-fixing and the ghost terms as well as the gauge invariant one; $S=S_{\rm SYM}+i\delta_{\rm BRST}(\bar{C}\!\cdot\! G)$. The dots in $J\!\cdot\!\phi$ and $\bar{C}\!\cdot\! G$ denote both the integration over the coordinates and the trace over the group indices. Under an infinitesimal change of the gauge function, $G\to G+\Delta G$, we have \begin{eqnarray} Z_{G+\Delta G}[J]-Z_G[J]\!\!\!&=&\!\!\!\int\!{\cal D}\phi \,\delta_{\rm BRST}(-i\bar{C}\!\cdot\!\Delta G)e^{-S+J\cdot\phi}{\nonumber}\\ \!\!\!&=&\!\!\!\int\!{\cal D}\phi\,(-i\bar{C}\!\cdot\!\Delta G) (J\!\cdot\!\delta_{\rm BRST}\phi)e^{-S+J\cdot\phi}, \end{eqnarray} where in the last equality we have used the BRST Ward-Takahashi identity. Therefore, the partition function with the gauge function $G+\Delta G$ is expressed as \begin{eqnarray} Z_{G+\Delta G}[J]\!\!\!&=&\!\!\!\int{\cal D}\phi \Bigl(1+(-i\bar{C}\!\cdot\!\Delta G)(J\!\cdot\!\delta_{\rm BRST}\phi)\Bigr) e^{-S+J\cdot\phi}{\nonumber}\\ \!\!\!&=&\!\!\!\int{\cal D}\phi \exp\left\{-S+J\!\cdot\!\left(\phi+i\delta_{\rm BRST}\phi\, (\bar{C}\!\cdot\!\Delta G)\right)\right\}. \label{ZG+DG} \end{eqnarray} This is nothing but the partition function for the field $\phi+i\delta_{\rm BRST}\phi\,(\bar{C}\!\cdot\!\Delta G)$ in the original gauge $G$. Our next task is to restate the property (\ref{ZG+DG}) in terms of the effective action $\Gamma_G[\varphi]$ which is related to $\ln Z_G[J]$ by the Legendre transformation. In appendix B, we show that (\ref{ZG+DG}) implies the following relation between $\Gamma_{G+\Delta G}$ and $\Gamma_G$: \begin{eqnarray} \Gamma_{G+\Delta G}[\varphi] =\Gamma_G[\varphi+\langle i\delta_{\rm BRST}\phi\, (\bar{C}\!\cdot\!\Delta G)\rangle_\varphi], \label{gaugeshift} \end{eqnarray} where $\langle{\cal O}\rangle_\varphi$ denotes the generating functional of the 1PI Green's function with an insertion of the operator ${\cal O}$ in the original gauge. Let us apply the general arguments given above to the relation between our new gauge (\ref{NEWgauge}) and the old one (\ref{BGgauge}) and reproduce the field redefinition (\ref{tildeB}) in the D-particle case. Using (\ref{gaugeshift}) with $\varphi=B_m$ and $\Delta G=\eta A$, and (\ref{freeprop}), we find that \begin{eqnarray} \widetilde B_m(\tau)\!\!\!&=&\!\!\! B_m(\tau)+\Bigl\langle[X_m,C]_{N+1,N+1}(\tau) \int\!{\rm d}\tau_0\mathop{\rm tr}\bar{C}\eta A(\tau_0)\Bigr\rangle{\nonumber}\\ \!\!\!&=&\!\!\! B_m-\int\!{\rm d}\tau_0\,\eta(\tau_0) \Bigl(\Bigl\langle C_{N+1,i}(\tau)\bar{C}_{i,N+1}(\tau_0)\Bigr\rangle \Bigl\langle Y_{m,i,N+1}(\tau)A_{N+1,i}(\tau_0)\Bigr\rangle -(i\leftrightarrow N+1)\Bigr){\nonumber}\\ \!\!\!&=&\!\!\! B_m-4{g_{\rm YM}^2} N\bra{\tau}\Delta\dot B_m\Delta\eta\Delta\ket{\tau} =B_m-{g_{\rm YM}^2} N\frac{3}{4}\frac{\eta\dot B_m}{B^5}, \label{tildeBagain} \end{eqnarray} where we have kept only those terms with the number of derivatives less than or equal to two. Note that $\Delta G=\eta A$ can be regarded as infinitesimal since we are interested only in the lowest non-trivial order in $\eta$. The result (\ref{tildeBagain}) agrees exactly with (\ref{tildeB}) proposed by \cite{JKY0}. \section{Matrix model in the $R_\xi$ gauge} In this section, we shall consider the Matrix model in another gauge. This will give a support for the validity of the general arguments of the previous section through a non-trivial calculation. It will also give a lesson about the structures of the field redefinitions and the non-renormalization theorem. The gauge we take here is the $R_\xi$ gauge with $|1-\xi|\ll 1$: \begin{eqnarray} G=-{\partial} A+i\xi\,[B_m,Y_m]+{1 \over 2}\xi{g_{\rm YM}^2} b, \end{eqnarray} which has the properties that for $\xi=1$ it reduces to the original background gauge, and that the $Y$-$A$ mixing is given by ${\cal L}_{YA}$ of (\ref{quadratic}) independently of the value of $\xi$. For an infinitesimal $\alpha\equiv1-\xi$, the difference from the original background gauge is \begin{equation} \Delta G=\alpha\Bigl(-i[B_m,Y_m]-{1 \over 2}{g_{\rm YM}^2} b\Bigr) =\frac{\alpha}{2}\Bigl(-{\partial} A+i[B_m,Y_m]\Bigr). \end{equation} {}From the general formula (\ref{gaugeshift}) of the previous section, we can obtain the field redefinition relating the $R_\xi$ gauge and the original background gauge ($\xi=1$). The calculation is carried out by using the free propagators for $\xi=1$, \begin{eqnarray} \langle C_{ij}(\tau_1)\bar{C}_{ji}(\tau_2)\rangle\!\!\!&=&\!\!\! i\bra{\tau_1}\Delta_{ij}\ket{\tau_2},{\nonumber}\\ \langle Y_{m,ji}(\tau_1)A_{ij}(\tau_2)\rangle\!\!\!&=&\!\!\! -2i\bra{\tau_1}\Delta^{\eta}_{ij} V_{m,ij}\Delta^{\eta}_{ij} \Bigl(1-4V_{n,ij}\Delta^{\eta}_{ij} V_{n,ij}\Delta^{\eta}_{ij}\Bigr)^{-1}{g_{\rm YM}^2}(\tau) \ket{\tau_2},{\nonumber}\\ \langle Y_{m,ij}(\tau_1)Y_{n,ji}(\tau_2)\rangle\!\!\!&=&\!\!\!\delta_{mn} \bra{\tau_1}\Delta^{\eta}_{ij} \Bigl(1-4V_{\ell,ij}\Delta^{\eta}_{ij}V_{\ell,ij}\Delta^{\eta}_{ij}\Bigr)^{-1} {g_{\rm YM}^2}(\tau)\ket{\tau_2},{\nonumber} \end{eqnarray} and the formulas given in the appendix A, and taking into account the derivatives of the coupling ${g_{\rm YM}^2}$ carefully. We find \begin{eqnarray} \widetilde B_m\!\!\!&=&\!\!\! B_m +\alpha{g_{\rm YM}^2} N \Biggl(\frac{1}{4}\frac{B_m}{B^3} -\frac{1}{16}\frac{\ddot B_m}{B^5} +\frac{5}{8}\frac{\dot B\!\cdot\! B\,\dot B_m}{B^7} -\frac{5}{16}\frac{\ddot B\!\cdot\! B\,B_m}{B^7}{\nonumber}\\ &&\quad -\frac{5}{16}\frac{\dot B^2B_m}{B^7} +\frac{35}{32}\frac{(\dot B\!\cdot\! B)^2B_m}{B^9} -\frac{5}{16}\frac{\eta\,\dot B_m}{B^5} -\frac{5}{8}\frac{\eta\,\dot B\!\cdot\! B\,B_m}{B^7} +\frac{3}{16}\frac{\dot\eta\,B_m}{B^5}\Biggr).\label{redefinition} \end{eqnarray} To confirm the field redefinition (\ref{redefinition}), let us next consider the 1-loop effective action in the $R_\xi$ gauge: \begin{eqnarray} \Gamma_{1\hbox{\scriptsize -loop}}\!\!\!&=&\!\!\! N\mathop{\rm Tr}\biggl\{ \ln\Bigl((-{\partial}^2+\eta{\partial}+B^2)\delta_{mn}-(1-\xi)B_mB_n\Bigr) +\ln\Bigl(-{\partial}^2+\eta{\partial}+\xi B^2\Bigr){\nonumber}\\ &&+\ln\biggl(1-4V_m\Bigl(\Delta^{\eta} B_mB_n\Delta^{\eta}\Bigr)V_n \xi\Bigl(-{\partial}^2+\eta{\partial}+\xi B^2\Bigr)^{-1}\biggr){\nonumber}\\ &&-2\ln\Bigl(-{\partial}^2+\xi B^2\Bigr) -4\sum_{\pm}\ln\Bigl(-{\partial}^2+B^2\pm\dot B\Bigr)\biggr\}, \label{GoneloopRxi} \end{eqnarray} where $\mathop{\rm Tr}$ for the first term implies also the trace operation with respect to $(m,n)$. Keeping only those terms proportional to $\alpha$ and at most linear in $\eta$, and further with number of derivatives less than or equal to four, we get after tedious but straightforward calculations the following result for the shift of the effective action in the $R_\xi$ gauge from that in the $\xi=1$ gauge: \begin{eqnarray} &&\int\!{\rm d}\tau\alpha N \biggl(\frac{1}{4}\frac{\dot B^2}{B^3} -\frac{3}{4}\frac{(\dot B\!\cdot\! B)^2}{B^5} -\frac{1}{16}\frac{\dddot B\!\cdot\!\dot B}{B^5} -\frac{5}{16}\frac{\dddot B\!\cdot\! B\,\dot B\!\cdot\! B}{B^7} -\frac{5}{8}\frac{\ddot B\!\cdot\!\dot B\,\dot B\!\cdot\! B}{B^7}{\nonumber}\\ &&\quad +\frac{35}{8}\frac{\ddot B\!\cdot\! B(\dot B\!\cdot\! B)^2}{B^9} +\frac{105}{32}\frac{\dot B^2(\dot B\!\cdot\! B)^2}{B^9} -\frac{315}{32}\frac{(\dot B\!\cdot\! B)^4}{B^{11}} +\frac{1}{4}\frac{\eta\,\dot B\!\cdot\! B}{B^3}{\nonumber}\\ &&\quad +\frac{3}{16}\frac{\eta\,\dddot B\!\cdot\! B}{B^5} +\frac{7}{16}\frac{\eta\,\ddot B\!\cdot\!\dot B}{B^5} -\frac{5}{8}\frac{\eta\,\ddot B\!\cdot\! B\,\dot B\!\cdot\! B}{B^7} +\frac{5}{16}\frac{\eta\,\dot B^2\dot B\!\cdot\! B}{B^7} +\frac{35}{32}\frac{\eta\,(\dot B\!\cdot\! B)^3}{B^9}\biggr). \label{shiftGamma} \end{eqnarray} In obtaining (\ref{shiftGamma}), we used the cyclicity of the trace to put all the terms coming from the expansion of (\ref{GoneloopRxi}) into the standard forms of $\eta B_m\Delta\cdots\Delta f(\tau)\Delta$ or $\eta{\partial}\Delta\cdots\Delta f(\tau)\Delta$, and applied the formulas in appendix A. Then, our next task is to determine the redefinition $B_m\to\widetilde{B}_m$ in such a way that the sum of the kinetic term $\dot{B}^2/2{g_{\rm YM}^2}$ and the shift of $\Gamma_{1\hbox{\scriptsize -loop}}$ (\ref{shiftGamma}) is identified as the kinetic term of the new field $\widetilde{B}_m$. The condition for the identification is apparently overdetermined. For example, the four-derivative terms independent of $\eta$ in (\ref{shiftGamma}) must correspond to the shift of $B_m$ by some two-derivative terms. There are 5 kinds of such terms in the shift of $B_m$; $\ddot B_m/B^5$, $\dot B_m\dot B\!\cdot\! B/B^7$, $\cdots$, $B_m(\dot B\!\cdot\! B)^2/B^9$. On the other hand, there are 11 kinds of four-derivative terms independent of $\eta$ in the 1-loop effective action; $\ddddot B\!\cdot\! B/B^5$, $\dddot B\!\cdot\!\dot B/B^5$, $\cdots$, $(\dot B\!\cdot\! B)^4/B^{11}$. Besides, there are 5 kinds of total derivative terms as the ambiguity of the effective action; ${\rm d}/{\rm d}\tau[\dddot B\!\cdot\! B/B^5]$, ${\rm d}/{\rm d}\tau[\ddot B\!\cdot\!\dot B/B^5]$, $\cdots$, ${\rm d}/{\rm d}\tau[(\dot B\!\cdot\! B)^3/B^9]$. Therefore, we have 11 equations with only 5+5 unknowns. However, there is a solution and it coincides with (\ref{redefinition}) obtained from the general formula (\ref{gaugeshift}). We can also calculate the quantum SCT in the $R_\xi$ gauge to check the consistency of the field redefinition (\ref{redefinition}). Note the expression for the quantum SCT in the $R_\xi$ gauge is not changed form (\ref{extraterm}), and we obtain \begin{eqnarray} \delta_{\rm SCT}^{\rm Q} B_m=-{g_{\rm YM}^2} N\biggl\{ \Bigl(\frac{3}{2}+\alpha\Bigr)\frac{\epsilon\,\dot B_m}{B^5} +\frac{5}{4}\alpha\frac{\epsilon\,\dot B\!\cdot\! B\,B_m}{B^7} \biggr\}. \label{qSCTRxi} \end{eqnarray} One can easily see that the two quantum SCTs, (\ref{qSCTRxi}) and (\ref{Dlt_B_final}) with $p=0$, are consistently related by the redefinition (\ref{redefinition}). Finally in this section we shall give some comments. First, although it might seem strange that there appear two-derivative terms (the first two terms) in (\ref{shiftGamma}), it does not contradict the non-renormalization theorem of \cite{SSP}. The statement of \cite{SSP} is that if we remove the $(\dot B\!\cdot\! B)^2/B^5$ term by a suitable coordinate transformations, the other term $\dot B^2/B^3$ will automatically disappear. Actually, the $B_m/B^3$ term in (\ref{redefinition}) removes the two terms in (\ref{shiftGamma}) simultaneously. Our second comment is on the meaning of the field redefinitions. The redefinition (\ref{tildeB}) is simply a field dependent shift of the world-volume coordinate $\tau$ and hence is interpretable as a change of reparametrization gauge in the Born-Infeld action. However, the redefinition (\ref{redefinition}) contains, besides the terms interpretable as the target space coordinate transformation (the $B_m/B^3$ term) and the world-volume reparametrization (terms containing $\dot B_m$), all kinds of terms with a given dimension. \section{Conclusions and further directions} In this paper, we proposed a new gauge which is useful for discussing the conformal symmetry in the Matrix model. The form of the quantum modified SCT depends crucially on the choice of gauge in SYM, and our new gauge reproduced exactly the SCT of the isometry. We also gave a general argument on the relation between the change of gauge and the field redefinition in the effective action. Then, we examined the $R_\xi$ gauge as an example and reconfirmed the special nature of the background gauge and our new gauge. Namely, the diagonal elements of the Higgs fields in SYM correspond directly to the target space coordinates in the Born-Infeld action in the static gauge only when we take the background gauge or our new gauge in SYM. We shall discuss some further directions of our work. First, the higher loop analysis of the Matrix model in our new gauge is an interesting subject. As the isometry of the near-horizon geometry of the D-brane solution determines the Born-Infeld action and we can derive the isometry directly from the 1-loop calculation in SYM in our new gauge, we expect that the quantum modified SCT in SYM is essentially 1-loop exact. The analysis of the quantum modified SCT in higher loops rather than that of the effective action would be a cleverer way to show the full agreement between the Matrix model and supergravity. Our new gauge would also be useful for analyzing more general multi-D-particle systems than the simple source-probe configuration of this paper. The new gauge we proposed in this paper may be important even conceptually. The fact that the signs of the $\eta{\partial}$ in the kinetic terms (\ref{newgaugequadratic}) are opposite between $Y_m$ (the coordinates perpendicular to the branes) and $A_\mu$ (the coordinates parallel to the branes) reminds us of the spacetime uncertainty principle proposed by \cite{Y}. This would be a clue for the understanding of the deep meaning of our new gauge. \vspace{.6cm} \noindent {\Large\bf Acknowledgments}\\[.2cm] We would like to thank to our colleagues at Kyoto University for various useful discussions. The work of H.\ H.\ is supported in part by Grant-in-Aid for Scientific Research from Ministry of Education, Science and Culture (\#09640346). The work of S.\ M.\ is supported in part by the Japan Society for the Promotion of Science under the Predoctoral Research Program. \vspace{1cm} \noindent {\Large\bf Appendices}
\section{Introduction} Although the nature of the superconducting pair wave function in high -$T_c$ cuprates is not yet known strong evidences of a major $\rm d_{x^2-y^2}$ symmetry exists \cite{Cox,Harlingen,Scalapino}. Experiments sensitive to the internal phase structure of the pair wave function reported a sign reversal of the order parameter supporting $d$ wave symmetry \cite{dwave}. Most recently from various experiments and theory it appears that the pairing symmetry of these family could be a mixed one like $\rm d_{x^2-y^2} + e^{i\theta}{\alpha}$ where $\alpha$ could be something in the $s$ wave family or $d_{xy}$. There were early questions from tunneling experiments regarding the pure d-wave symmetry\cite{Dynes} as the data supports an admixture of d and s-wave components due to orthorhombicity in YBCO\cite{Walker,Carbotte}. Possibility of a minor but finite $id_{xy}$ symmetry alongwith the predominant $d_{x^2-y^2}$ has also been suggested\cite{SBL} in connection with magnetic defects or small fractions of a flux quantum $\Phi_0=hc/2e$ in YBCO powders. Similar proposals came from various other authors in the context of magnetic field, magnetic impurity, interface effect etc. \cite{Laughlin,Krishana,others} The experimental result by Krishana {\it et al.}, was interpreted as a signature of induction of a minor component {\it eg}, $id_{xy}$ or $is$ in a $d$-wave superconducture with the application of magnetic field along the $c$ axis. In this work, we study in details the effect of a weak Zeeman magnetic field on superconductors with mixed order parameter symmetry like $\Delta (k) = \Delta_{d_{x^2-y^2}} + e^{i\theta}\alpha$ with $\alpha=d_{xy}$, $s$ for arbitrary $\theta$. It is well known that such superconductors with $\theta \neq 0$ and $\alpha \neq 0$ corresponds to broken time reversal states (BTRS). These BTRS states lift the directional degeneracy of charge currents by admixing a subdominant $\alpha$-wave component to the d-wave pairing state and a spontaneous finite current appears\cite{trsbreview}. An application of Zeeman magnetic field can lift the spin degeneracy leading to suppression of BTRS ; a pure $d$-wave occurs with increasing magnetic field. For mixed symmetry as above with $\theta =0$ that preserves time reversal symmetry and are nodeful respond differently to the Zeeman field as compared to $\theta \neq 0$ states. For nodefull $\theta =0$ state, the local gap $\Delta (k)$ of small magnitude over the Fermi surface may be destructed with the application of Zeeman field leading to a paramagnet pocket. This although true for $\theta \neq 0$ states, but such states correspond to {\em fully gapped} situation all over the Fermi surface which causes {\em weak} response to the magnetic field. A clear picture on the above will be demonstrated in this article. It may be mentioned that the high temperature superconductors are quasi two dimensional in nature and therefore, a magnetic field parallel to the $\rm Cu-O$ plane does not couple to the orbital motion of the electrons in the plane. Therefore, we shall not consider spin-orbit interaction in this work. In connection with the discussion of order parameter symmetry in cuprates, we would further like to mention that the proposal of mixed order parameter symmetry got the correct momentum when experimental data on longitudinal thermal conductivity by Krishana {\it et al,} \cite{Krishana} of $\rm Bi_2Sr_2CaCu_2O_8$ compounds and that by Movshovich {\it et al,} \cite{Krishana} showed supportive indication to such proposals. There are experimental results related to interface effects as well as in the bulk that indicates mixed pairing symmetry (with dominant $d$-wave) \cite{others}, thus providing a strong threat to the pure $d$ wave models. There were early orginal works as regards to the modifaction of superconductivity due to application of Zeeman magnetic \cite{various} and very recently, the Zeeman suppression was discussed in mesoscopic systems \cite{braun}. \section{Model Calculation} The free energy of a two dimensional planar superconductor with arbritary pairing symmetry in presence of a magnetic field may be written as, \begin{equation} F_{k,k^\prime}(h) = -\frac{1}{\beta} \sum_{k,\sigma = \pm} \ln (1 + e^{-\sigma \beta E_{k}^\sigma}) + \frac{\mid \Delta_k \mid^2}{V_{k k^\prime}} \label{free} \end{equation} where $E_{k}^{\sigma} = \sqrt{(\epsilon_{k}-\mu)^2 + \mid \Delta_k \mid^2} + \sigma (g\mu_B/2)B$ are the energy eigen values of a Hamiltonian that describes superconductivity, $(g\mu_B/2)$ is the magnetic moment of the electrons. This includes the assumption that the Zeeman field raises/lowers the energy of the spin up/down quasiparticle states. We minimize the free energy, Eq.\ (\ref{free}) {\it i.e}, $\partial F/\partial \mid \Delta\mid$ = 0, to get the gap equation as, \begin{equation} \Delta_k = \sum_{k^\prime} \frac{V_{kk^\prime}}{2} \frac{\Delta_{k^\prime}} {2 E_{k^\prime}}\left( \tanh (\frac{\beta E_{k^\prime}^+}{2})+ \tanh (\frac{\beta E_{k^\prime}^-}{2})\right) \label{gapeq} \end{equation} where $\epsilon_k$ is the dispersion relation taken from the ARPES data \cite{Ding} and $\mu$ the chemical potential will control band filling through a number conserving equation given below. Since the applied Zeeman field modifies the SC quasiparticles of spin up and down differently, their occupation probablities are also modified. The number conserving equation that controls the band filling through chemical potential, $\mu$ in presence of Zeeman field is given by, \begin{eqnarray} &&\rho(\mu,T,h)= \nonumber \\ && \sum_k\left[1-\frac{1}{2}\frac{(\epsilon_k - \mu)}{E_k}\left( \tanh\frac{\beta E_{k}^+}{2} + \tanh\frac{\beta E_{k}^-}{2}\right) \right] \label{dens} \end{eqnarray} where $h=(g\mu_B/2)B$. Let us consider that the overlap of orbitals in different unit cells is small compared to the diagonal overlap. Then in the spirit of tight binding lattice description, the matrix element of the pair potential used in the SC gap equation Eq. (\ref{gapeq}) may be obtained as, \begin{eqnarray} V(\vec q) &=& \sum_{\vec \delta} V_{\vec \delta} e^{i \vec q \vec R_{\delta}} = V_{0} + V_1 f^d (k) f^d(k^\prime) + V_1 g(k) g(k^\prime) \nonumber \\ && + V_2 f^{d_{xy}}(k)f^{d_{xy}}(k^\prime) +V_2 f^{s_{xy}}(k)f^{s_{xy}}(k^\prime) \label{ppot} \end{eqnarray} where in the first result of the equation \ (\ref{ppot}) $\vec R_{\delta}$ locates nearest neighbour and further neighbours, $\vec \delta$ labels and $V_n$, $n=1,2$ represents strength of attraction between the respective neighbour interaction. The first term in the above equation $V_{0}$ refers to the on-site interaction which has an effective attractive value giving rise isotropic $s$ wave. The second and third terms are responsible for $d$ and extended $s$ wave symmetry superconductivity whereas the $4^{th}$ and the $5^{th}$ terms are responsible for $d_{xy}$ and $s_{xy}$ symmetries respectively. We restrict only to singlet pairing states ({\it i.e}, $\Delta (k)= \Delta (-k)$) as applicable for high temperature superconductors. The momentum form factors are obtained as, \begin{eqnarray} & &f^d(k) = \cos (k_xa) - \cos (k_ya) \nonumber \\ && g(k) = \cos (k_xa) + \cos (k_ya) \nonumber \\ && f^{d_{xy}}(k)= 2 \sin (k_xa) \sin (k_ya) \nonumber \\ && f^{s_{xy}}(k)= 2 \cos (k_xa) \cos (k_ya) \label{symm} \end{eqnarray} For two component order parameter symmetries as mentioned above, we substitute the required form of the potential and the corresponding gap structure into the either side of Eq. \ (\ref{gapeq}) which gives us an identity equation. Then separating the real and imaginary parts together with comparing the momentum dependences on either side of it we get gap equations for the amplitudes in different channels as, \begin{eqnarray} \Delta_j= \sum_k\frac{V_j}{2}\frac{\Delta_jf^{j^2}_{k}}{2E_k}\left[\tanh( \frac{\beta E_{k}^+}{2})+\tanh(\frac{\beta E_{k}^-}{2}) \right] \label{gapcomp} \end{eqnarray} where $j = 1, 2$ corresponding to two components $d$ and $s$ or $d$ and $d_{xy}$ symmetries. Considering mixed symmetry of the form $\Delta (k) = \Delta_{d_{x^2-y^2}}(0)f^d (k) +e^{i\theta}\Delta_{s}(0)$ one identifies $\Delta_1 = \Delta_{d_{x^2-y^2}}(0)$, $\Delta_2 = \Delta_{s}(0)$, $f_{k}^1 = f^d (k)$, $f_{k}^2 = 1$ and $V^1=V_1$, $V^2=V_0$ in Eq. (4). Similarly, for mixed symmetries of the form $\Delta (k) = \Delta_{d_{x^2-y^2}}(0)f^d (k) +e^{i\theta}\Delta_{d_{xy}}(0)f^{d_{xy}}$ $\Delta_2 = \Delta_{d_{xy}}(0)$, $f_{k}^2=f_{k}^{d_{xy}}$ and $V^1=V_1$, $V^2=V_2$ of Eq.(4). The potential required to get such pairing symmetries are discussed in Eq. (\ref{symm}). We solve self-consistently the above three equations (Eq.\ref{gapcomp} and Eq.\ref{dens}) in order to study the phase diagram of a mixed order parameter superconducting phase in presence of Zeeman magnetic field. The numerical results obtained for the gap amplitudes through Eqs. \ (\ref{gapcomp},\ref{dens}) will be compared with free energy minimizations via Eq. \ (\ref{free}) to get the phase diagrams. \section{Results and Discussions} We present in this section our numerical results for a set of fixed parameters, {\it e.g}, a cut-off energy $\Omega_c$= 500 K around the Fermi level above which superconducting condensate does not exist, a fixed transition temperature of the minor component $T_{c}^\alpha (h=0) = 24$ K and the bulk $T_c = 85$ K determined by the $d$-wave order parameter. In figures 1 and 2 we present results for $\Delta (k) = \Delta_{d_{x^2-y^2}}(0)f^d (k) + e^{i\theta}\alpha$ symmetries for $\theta = \pi/2$ and $\theta =0$ respectively. Such symmetries would arise from a combination of two component pair potentials ($2^{nd},~1^{st}$), ($2^{nd},~4^{th}$) terms in Eq. \ref{ppot} for $\alpha= s ~\& ~d_{xy}$ respectively. The amplitudes of extended $s$-wave states like $s_{x^2+y^2}$, $s_{xy}$ are found to be finite only towards very low band filling, $\rho \sim 0$ for $\theta=\pi/2$ hence does not cause any mixing with the predominant $d$-wave. Therefore, we shall discuss only the results of $\theta =0$ and $\theta =\pi/2$ for $\alpha= s ~\& ~d_{xy}$. These two phases of $\theta$ can cause important differences. It is known that for any $\theta \neq 0$, time reversal symmetry is locally broken \cite{hng2} at lower temperatures with the onset of the secondary component which correspond to a phase transition to an fully gapped phase for $h=0$, from a partially ungapped phase of $d_{x^2-y^2}$ symmetry. On the other hand, the $\theta=0$ phase still remains nodeful, although the nodal lines shifts a lot from the usual $k_x=k_y$ lines of the $d_{x^2-y^2}$. In Fig. 1(a) and (b) we present the temperature dependencies of the order parameters in the complex mixed symmetry. Curves corresponding to the $d$-wave channels are represented with thinner joining lines of different styles whereas the minor $s$ component is denoted through that of same style but with thicker lines (same strategy will be carried out in other figures as well). For a zero Zeeman field ($h=0$), the amplitude of the $d$-wave is suppressed with the onset of the minor $s$ component at $T=24$ K which leads to a kink like structure (cf. the thin solid curve in Fig1 (a)). \begin{figure} \epsfxsize=4.5truein \epsfysize=4.5truein {\epsffile{fig1.eps}} \caption{{\bf (a)} Amplitudes of the $\rm \Delta_{d_{x^2-y^2}}$ and $\rm \Delta_{s}$ at a fixed band filling $\rho = 0.85$ as a function of temperature (in K) for $\rm \theta = \pi/2$ ({\it i.e}, $\rm d_{x^2-y^2}+is$) phase in various values of Zeeman field ($h$). With the application of weak magnetic field while the $\rm d_{x^2-y^2}$ remains almost unaffected the transition temperature ($T_{c}^\alpha$) of the minor $\alpha =s$ component is suppressed strongly. At a field value of $h = 0.014$ eV the $s$-wave shows a first order transition. {\bf (b)} For field values $h \geq 0.016$ eV the minor $s$-channel is completely suppressed leading to a pure $d$ wave order parameter. The thermal behaviour of $d$-wave superconductivity in presence of the Zeeman field is presented in Fig.1(b). While the amplitudes at $T=0$ K remains almost unaffected with field, the transition temperatures get affected. A magnetic field induced first order transition is observed at $h = 0.04$ eV.} \label{fig:dsstar-complex} \end{figure} With application of weak Zeeman field the transition temperature of the minor $s$ state decreases while the zero temperature magnitude remains the same. This causes a shift in the kink like structure in the thermal dependence of the $d$ wave channel towards lower temperature with increasing field and hence a small enhancement in the $d$-wave with field at lower temperature occurs. This point will be clearer from Fig. 5(b) as discussed latter. At a field value $h_c = 0.016$ eV, the $s$ wave component is completely suppressed leading to a pure $d$ wave phase. Thus we have a magnetic field induced transition at lower temperature from {\em fully gapped phase} of the $d+is$ state to a {\em partially} gapped phase of the $d$-wave. Therefore, in absence or very low magnetic field, there is a transition from a {\em partially gapped} phase of the $d$-wave at higher temperature to a {\em fully gapped} $d+is$ phase at lower temperatures. With increasing field the fully gapped phase region with respect to temperature decreases and brings back the ungapped phase of the $d$-wave. These phase transitions will have important bearings in the thermodynamic and transport properties. \begin{figure} \epsfxsize=3.25truein \epsfysize=3.25truein {\epsffile{fig2.eps}} \caption{Same as that of figure 1(a) except $\rm \theta = 0$ ({\it i.e,} $\rm d_{x^2-y^2} + s$ symmetry). For the $d$-wave channel, its amplitude is found to increase below $T_{c}^s$ in contrast to $\theta = \pi/2$ case where the $d$-wave amplitude is suppressed. The $s$-wwave channel suffers drastic suprresion in $T_{c}^s$ as well as the zero temperature amplitude in contrast to that in Fig. 1(a). Note, the critical field ($h_c$) at which the $s$ component is completely suppressed is 0.013 eV in contrast to 0.016 eV in case of $\theta = \pi/2$ (cf. Fig1a).} \label{fig:dsstar-real} \end{figure} In Fig. 1(b) we present the paramagnetic state of the $d$-wave. This phase has also been qualitatively investigated recently by Yang and Sondhi \cite{yang} with possibility of pairing with finite momentum. We therefore restrict to present only the details thermal dependence of the $d$-wave superconductivity in presence of Zeeman magnetic field that were not discussed. We show that with increasing field (at lower fields) the zero temperature magnitude of the $d$-wave remains unchanged whereas the $T_c$ is reduced. At higher field {\it e.g}, $h=0.04$ one sees a first order transition from superconducting state to the normal state with respect to temperature. This behaviour causes a magnetic field induced enhancement of the $2 \Delta/k_BT_c$ ratio. This ratio is crucial for many physical properties like specific heat jump etc. and hence expected to have drastic effect with magnetic field. \begin{figure} \center \epsfxsize=3.25truein \epsfysize=3.25truein \leavevmode {\epsffile{fig3.eps}} \caption{Amplitudes of the $\rm \Delta_{d_{x^2-y^2}}$ and $\rm \Delta_{d_{xy}}$ at a fixed band filling $\rho = 0.85$ as a function of temperature (in K) for $\rm \theta = \pi/2$ ({\it i.e}, $\rm d_{x^2-y^2}+id_{xy}$ in various values of Zeeman field ($h$). With the application of weak magnetic field while the $\rm d_{x^2-y^2}$ remains practically unaffected the transition temperature ($T_{c}^\alpha$) of the minor $\alpha =d_{xy}$ component is suppressed strongly. Similar to that of Fig. 1a, a first order transition is seen in the minor $d_{xy}$ channel for $h=0.014$ eV.} \label{fig:dsxy-complex} \end{figure} In Fig.2 we describe the thermal behaviour of superconductors with $d+s$ symmetry in presence of magnetic field. First of all, in absence of magnetic field, the thermal behaviors at lower temperatures is quite different (as we have seen in Fig. 1), the $s$-wave gap opens very fast below $T=24$ K and also induces a growth to the $d$ at a faster rate than that above $T=24$ K. While the $s$-wave has about three times the zero temperature value compared to $d+is$ phase, the $d$ wave also have quite larger value. With application of a very small magnetic field the $s$-component is suppressed largely; both its $T_c$ and the zero temperature gap being suppressed. The $d$-wave gap magnitude is also suppressed while its $T_c$ remained practically unchanged with such small values of the magnetic field. More importantly, although both the $d$ and the $s$ channel has larger magnitude in the $d+s$ phase the critical field ($h_c =0.013$ eV) at which the $s$ component vanishes completely is {\em smaller} compared to that for the $\theta = \pi/2$ phase ($h_c =0.016$ eV) of the mixed symmetry. Distinctly, the response of the Zeeman field to the $d+s$ superconductors is more pronounced than the $d+is$ superconductors. This study therefore also revealed the importance of the phase of the minor component in $d$-wave superconductors with a pratical example of effect of Zeeman magnetic field, for the first time. In order to establish the importance of the phase of the minor component we also study the effect of Zeeman magnetic field in $\rm d_{x^2-y^2}+id_{xy}$ and $\rm d_{x^2-y^2}+d_{xy}$ symmetry superconductors respectively. The qualitative as well as quantitative behaviors remain almost same as that shown in figures 1 and 2. Thus establishing that the response of the superconducors with mixed order parameter symmetries with $\theta=0$ phase of the minor component is stronger. \begin{figure} \epsfxsize=3.25truein \epsfysize=3.25truein {\epsffile{fig4.eps}} \caption{ Same as that in figure 3 except for $\rm \theta=0$ {\it i.e} $\rm d_{x^2-y^2}+d_{xy}$ phase that preserves the time reversal symmetry. The notable difference is that the minor $\rm d_{xy}$ component although have a magnitude at $T=0$ K and $h = 0$ three times larger than that for $\theta = \pi/2$ case, the minor component is suppressed at a lower critical field $h_c = 0.013$ eV.} \label{fig:dsxy-real} \end{figure} In figures 5 and 6 we concentrate on studying behavior of such mixed order parameter symmetry superconductors in presence of magnetic field as a function of band filling ($\rho$) for $d_{xy}$ and $s$ as minor components respectively. With application of a small field the minor component $\alpha=d_{xy}$ or $s$ is suppressed only at lower fillings and then with increasing field it is suppressed in the optimal doping regime suddenly. Therefore, with increasing field one finds mixed symmetry region as well as pure $d$-wave region with respect to filling \--- a {\em non-uniform} superconductivity. This behavior depends on the nature of the minor component. For $\alpha =s$, the mixed symmetry is possible around half-filling and at around $\rho=0.7$ in an intermediate field value. With increasing field mixed symmetry regions around both the band fillings shrinks and at a field value of $h=0.03$ the $s$-wave vanishes leading to a pure $d$-wave phase. It may be noticed the $d$-wave channel is enhanced (cf. Fig 5(b)) in the optimal doping region with intermediate field values as was also mentioned earlier while discussing Fig1 (a). This however does not occur in case of $d_{xy}$. For $d_{xy}$ minor component, the minor component is suppressed strongly only from the lower filling with increasing field and at an intermidiate field the mixing is possible only near half filling. The $d$-wave boundary towards lower fillings as well as around half-filling also shrinks with increasing field values in contrast to $\alpha=s$. One always see a sharp transition from a mixed phase to a pure $d$-wave phase or otherway around with respect to band filling irrespective of the minor component in a given magnetic field. At larger fields ($h > 0.03$) when the minor component is suppressed completely {\it i.e}, one has a pure $d$-wave phase, the $d$-wave phase also show sharp transition from $d$-wave superconducting state to a normal state (cf. Fig. 6). \begin{figure} \epsfxsize=4.5truein \epsfysize=4.5truein \leavevmode {\epsffile{fig5.eps}} \caption{Zero temperature amplitudes of the different pairing channels as a function of band fillings in the $\rm d_{x^2-y^2} + e^{i\theta}{\alpha}$ picture for several fields with $\theta = \pi/2$. An increase in the amplitude of the $d$-wave channel with some particular field values in the $\rm d_{x^2-y^2} + is$ picture near optimal doping is worth noticing. With field increasing the order parameter symmetry changes to pure $d$-wave at optimal and larger doping. } \label{fig:dsxy.temp} \end{figure} In Fig. 6 we study the same as that in Fig. 5 for $d+s$ and $d+d_{xy}$ symmetries. In contrast to the $\theta =\pi/2$ phase of the minor component (as in Fig. 5), the minor components have very large values in the $\theta =0$ phase although have the same $T_{c}^\alpha$ as mentioned earlier, in absence of the magnetic field. With very small magnetic field such large amplitudes of the condensation in the minor channel is strongly suppressed (see Fig 6(b) specially). The nature of suppression, as in Fig.5, is different for different minor component. For example, $d_{xy}$ is suppressed only from the lower filling whereas the $s$-wave is suppressed both from the lower filling as well as around the optimal doping. \begin{figure} \center \epsfxsize=4.5truein \epsfysize=4.5truein \leavevmode {\epsffile{fig6.eps}} \caption{Same as figure 5 except $\theta = 0$. Stronger and rapid suppression of the minor channel with field in contrast to that in Fig. 3. A visible change in the dominant $d$-wave channel is also seen in contrast to that in Fig. 3. Note, in case of $d+s$ picture, {\bf (b)}, the $d$-wave channel shows plateau with respect to band filling.} \label{fig:dsstar.temp} \end{figure} Overall, it is very distinct that the magentic field affects the $\theta =0$ mixed phase more stongly than the $\theta =\pi/2$ phase. This is due to the fact that the response of applied Zeeman field is paramagnetic with destruction of superconductivity over parts of the Fermi surface where the Zeeman field exceeds the local magnitude of the $k$-dependent gap resulting a spin polarization in the normal electrons. For $\theta = \pi/2$ phase, the nodes are missing and the Fermi surface is gapped all over, although the gap will have local minima. Thus response of the $\theta =\pi/2$ phase is weaker compared to the $\theta = 0$ phase. Noticiably, the critical value at which the minor component vanishes completely in all band fillings for the $\theta =0$ phase is $h_c = 0.025$ eV whereas that for $\theta =\pi/2$ is $h_c = 0.03$ eV. (The value of $h_c$ depends on $\rho$, as well as $\alpha$ and $\theta$). Above $h_c=0.025$, the paramegnetic state of the $d$-wave superconductors are also presented. The region of $d$-wave superconductivity shrinks with increasing field around the band filling $\rho = 0.82$. The change from $d$-wave to spin polarized normal state is very sharp with respect to the band filling. In the $d+s$ state a plateau has been observed in the $d$ channel in weak or zero field, similar to the behaviour known for the YBCO systems \cite{19}. It may be mentioned that Krishana {\it et al.}, found the longitudinal thermal conductivity of $\rm Bi_2Sr_2CaCu_2O_{8+\delta}$ at lower temperatures (5K to 20 K) decreases with the increase in magnetic field applied along $\rm c$-axis. Above a critical value of the magnetic field $\rm H_k(T)$, the thermal conductivity cease to change with the magnetic field and develops a plateau. It was proposed \cite{Krishana} that the $\rm d_{x^2-y^2}$ pairing state is unstable against the formation of $\rm d + e^{i \theta}\alpha$ (where $\rm \alpha = s, d_{xy}$) in presence of $\rm H_k (T)$ such that the loss of quasiparticle transport in the thermal conductivity can be explained. In contrast, we started with a $\rm d + e^{i \theta}\alpha$ picture and application of Zeeman field (which may be mapped as application of magnetic field parallel to the 2D $\rm Cu-O$ plane) without considering the orbital effect did not find any enhancement in the condensation of the minor channel but suppression leading to a pure paramegnetic $d$ state. \section{Summary} In summary we have performed a detailed study on the effect of Zeeman magnetic field on mixed pairing symmetry with predominant $d$-wave which seems to be very promising symmetry for the high $T_c$ systems. Thus we have described the paramagnetic state in the mixed symmetry superconductors and subsequently in the $d$-wave superconductors. In particular we established that the phase of the minor component mixed with predominant $d$-wave is of immense importance. The $\theta =0$ phase minor component symmetry responds to Zeeman field more profoundly than the $\theta =\pi/2$ of the minor component. It will be very interesting to calculte the specific heat, Magnetization, density of states as a function of magnetic field using this model. We argued that the orbital effect is secondary when a magnetic field is applied parallel to the conducting plane. This may indicate that the experimental observation by Krishana {\it et al.}, involve strong coupling of spins to orbitals due to application of magnetic field perpendicular to the plane at lower temperatures. However, the order parameter alone does not completely determine the thermodynamic property of a system. In Ref. \cite{yang}, some estimations and scaling relations in the change of various physical properties due to Zeeman field are given for a pure $d$-wave superconductor. It turns out, a weak Zeeman field does little to the order parameter but may profoundly affect the thermodynamic property of a pure $d_{x^2-y^2}$ supercondcutor. Such effects will presumably also remain for mixed superconductors with $\theta = 0$ as the gapnodes similar to $d_{x^2-y^2}$ remains, its effect in the $\theta =\pi/2$ phase has to be studied more carefully. \section{Acknowledgments} A large part of this work was carried out at the Instituto de F\'isica, Universidade Federal Fluminense, Brazil and was financially supported by the Brazilian funding agency FAPERJ, project no. E-26/150.925/96-BOLSA.
\section{Formulation of the model} High temperature superconductivity in the lanthanum \cite{bednorz1}, yttrium \cite{wu1} and related copper-oxide compounds remains a subject of intensive investigation and controversy. It was suggested that electron-phonon interaction mechanism, which is very successful in understanding of conventional (``low temperature") superconductors within the Bardeen-Cooper-Schriffer scheme \cite{bardeen1}, may not be adequate for high-$T_c$ cuprates, and even the conventional Fermi liquid model of metallic state may require reconsideration. This opens an area for investigation of mechanisms of electron-electron interaction which can be relevant in understanding peculiarities of superconducting, as well as normal state, properties of cuprates. Specific to all of them is the existence of oxide orbitals. Band calculations \cite{hybertsen1,mahan1} suggest that hopping between the oxygen $p_x,p_y$ orbitals and between the copper $d_{x^2-y^2}$ orbitals may be of comparable magnitude. On the experimental side, spectroscopic studies \cite{nucker1,kuiper1} clearly show that the oxygen band appears in the same region of oxygen concentration in which superconductivity in cuprates is the strongest. Therefore there exists a possibility that specific features of oxide compounds may be related to oxygen-oxygen hopping, or to the interaction between the copper and the rotational $p_x-p_y$ collective modes. If the oxygen hopping is significant then it immediately follows that intrinsic oxygen carriers ($p_x,p_y$ oxygen holes) should be different from the more familiar generic $s$-orbital derived itinerant carriers. The difference is related to low atomic number of oxygen such that removing or adding of one electron to atom induces a substantial change in the Coulomb field near the remaining ion and therefore results in the change of the effective radius of atomic orbitals near the ion. This will strongly influence the hopping amplitude between this atom and the atoms in its neighborhood. Such ``orbital contraction'' effect represents a source of strong interaction which does not simply reduce to the Coulomb (or phonon) repulsion (or attraction) between the charge carriers. It was suggested by Hirsch and coauthors \cite{hirsch1,hirsch2,hirsch3}, and by the present authors \cite{kulik1,kulik2,boyaci1,kulik3} that the occupation dependent hopping can have relevance to the appearance of superconductivity in high-temperature oxide compounds. In the present paper, we investigate the generic occupation-dependent hopping Hamiltonians with respect to peculiarities of the normal state, and to the range of existence of the superconducting state. Theoretical investigation of Cooper instability is supplemented by numeric study of pairing and diamagnetic currents in finite atomic clusters. We study the effect of Cooper pairing between the carriers and show that at certain values and magnitudes of the appropriate coupling parameters, the system {\em is} actually superconducting. The properties of such superconducting state are in fact only slightly different from the properties of conventional (low-$T_c$) superconductors. Among those we so far can only mention the change in the fluctuation conductivity above or near the critical temperature $T_c$. Relaxation of the pairing parameter to equilibrium acquires a small real part due to the asymmetry of contraction-derived interaction between the quasi-particles above and below the Fermi energy. Oxygen atoms in the copper-oxygen layers of the cuprates (Figure 1) have simple quadratic lattice. We assume that $p_z$ orbitals of oxygen ($z$ is the direction perpendicular to the cuprate plane) are bound to the near cuprate layers whereas carriers at the $p_x,p_y$ orbitals may hop between the oxygen ions in the plane. Let $t_1$ be the hopping amplitude of $p_x \, (p_y)$ and $t_2$ the hopping amplitude of $p_y \, (p_x)$ oxygen orbitals between the nearest lattice sites in the $x \, (y)$ direction in a square lattice with a lattice parameter $a$. Then the non-interacting Hamiltonian is \begin{equation} H_0=-t_1 \sum_{<ij>_x} a_i^{\dagger}a_j - t_2 \sum_{<ij>_y} a_i^{\dagger}a_j -t_1 \sum_{<ij>_y} b_i^{\dagger} b_j - t_2 \sum_{<ij>_x}b_i^{\dagger} b_j \end{equation} where $a_i^{\dagger} \, (a_i)$ is the creation (annihilation) operator for $p_x$ and correspondingly $b_i^{\dagger} \, (b_i)$ for $p_y$ orbitals. The interaction Hamiltonian includes the terms \begin{equation} H_1=\sum_{<ij>} a_i^{\dagger}a_j \left[ V m_i m_j + W (m_i+m_j) \right] + \sum_{<ij>} b_i^{\dagger}b_j \left[V n_i n_j + W (n_i +n_j) \right] \end{equation} where $n_i=a_i^{\dagger}a_i$, $m_i=b_i^{\dagger}b_i$. This corresponds to the dependence of the hopping amplitude on the occupation numbers $n_i, \, m_i$ of the form \begin{equation} \left( \hat{t}_{ij} \right)_{a_i \rightarrow a_j} = \tau _0 (1-m_i)(1-m_j) +\tau _1 \left[ (1-m_i)m_j+m_i(1-m_j) \right] + \tau_2 m_i m_j \end{equation} and correspondingly $(\hat{t}_{ij})_{b_i \rightarrow b_j}$ of the same form with $m_i$ replaced with $n_i$. The amplitudes $\tau_0, \, \tau_1, \, \tau_2$ correspond to the transitions between the ionic configurations of oxygen: \begin{eqnarray} \tau_0: \, & O_i^-+O_j^{2-} & \rightarrow O_i^{2-}+O_j^- \nonumber\\ \tau_1: \, & O_i+O_j^{2-} & \rightarrow O_i^{2-}+O_j \\ \tau_2: \, & O_i+O_j^- & \rightarrow O_i^- +O_j \nonumber \end{eqnarray} $O$ corresponds to the neutral oxygen ion whereas $O^-$ to the single charged and $O^{2-}$ to the double charged negative ions. Since oxygen atom has $1s^22p^42s^2$ configuration in its ground state, filling of the $p$ shell to the full occupied configuration $2p^6$ is the most favorable. Amplitudes $V$ and $W$ relate to the parameter $\tau_0, \, \tau_1, \, \tau_2$ according to \begin{equation} V=\tau_0-2\tau_1+\tau_2, \: W=\tau_1-\tau_2 \, . \end{equation} Assuming $t_1=t_2$ and replacing $a_i, \, b_i$ with $a_i$ with the pseudo-spin indices $\sigma=\downarrow , \, \uparrow$ we write the Hamiltonian Eq.(1) in the form \begin{equation} H=-t\sum_{<ij> \, {\sigma}} a_{i \, \sigma}^{\dagger}a_{j \, \sigma} +H_U+H_V+H_W \end{equation} where \begin{eqnarray} H_U &=&U \sum_i n_{i \uparrow} n_{i \downarrow} \\ H_V &=& V \sum_{<ij> \, \sigma} a_{i \sigma}^{\dagger} a_{j \sigma} n_{i, \bar{\sigma}} n_{j, \bar{\sigma}} \\ H_W &=& W \sum_{<ij> \, \sigma} a_{i \sigma}^{\dagger} a_{j \sigma} (n_{i, \bar{\sigma}} +n_{j, \bar{\sigma}}) \end{eqnarray} where we also included the in-site Coulomb interaction (U) between the dissimilar orbitals at the same site. $\sigma$ can also be considered as a real spin projection of electrons at the site. In that case, the pairing will originate between the spin-up and spin-down orbitals, rather than between $p_x$ and $p_y$ orbitals. More complex mixed spin -and orbital- pairing configurations can also be possible within the same idea of orbital contraction (or expansion) at hole localization but are not considered in this paper. The following discussion does not distinguish between the real spin and the pseudo-spin pairing. The Hamiltonian, Eq.(6), is a model one which can not refer to the reliable values of the parameters appropriate to the oxide materials. The purpose of our study is rather to investigate the properties of superconducting transition specific to the model chosen and to find the range of the $U, \, V, \, W$ values which may correspond to superconductivity. This will be done along the lines of the standard BCS model \cite{abrikosov1} in the weak coupling limit, $U, \, V, \, W \, \rightarrow \, 0$, and by an exact diagonalization of the Hamiltonian for a finite atomic cluster at large and intermediate coupling. In the momentum representation, the Hamiltonian becomes $H=H_0+H_1+H_2$ with \begin{eqnarray} H_0=\sum_{{\bf p} \, \sigma} \xi_{\bf p} a_{{\bf p} \sigma}^{\dagger} a_{{\bf p} \sigma}\\ H_1=\frac{1}{4} \sum_{p_1 p_2 p_3 p_4, \alpha \beta \, \gamma \, \delta} a_{{\bf p}_1 \alpha}^{\dagger} a_{{\bf p}_2 \beta}^{\dagger} \Gamma_{\alpha \beta \gamma \delta}^0(p_1, p_2 ,p_3 ,p_4) a_{{\bf p}_4 \delta} a_{{\bf p}_3 \gamma} \end{eqnarray} where \begin{equation} \xi_{\bf p} = -t\sigma_{\bf p}-\mu, \;\;\: \sigma_{\bf p}=2(\cos p_xa+ \cos p_y a), \end{equation} and $\mu$ is the chemical potential. $\Gamma_{\alpha \beta \gamma \delta}^0$ is the zero order vertex part defined as \begin{equation} \Gamma_{\alpha \beta \gamma \delta}^0(p_1, p_2, p_3, p_4)= \left[ U+(W+ \frac{1}{2}\nu V) (\sigma_{{\bf p}_1} +\sigma_{{\bf p}_2}+\sigma_{{\bf p}_3}+\sigma_{{\bf p}_4}) \right] \tau_{\alpha \beta}^x \tau_{\gamma \delta}^x(\delta_{\alpha \gamma} \delta_{\beta \delta} - \delta_{\alpha \delta} \delta_{\beta \gamma}) \delta_{{\bf p}_1+{\bf p}_2, {\bf p}_3+{\bf p}_4} \end{equation} where $\tau_{\alpha \beta}^x$ is a Pauli matrix \begin{eqnarray*} \left( \begin{array}{lr} 0 & 1 \\ 1 & 0 \end{array} \right) . \end{eqnarray*} For reasons which will be clear later, we separated $H_V$ and put some part of it into the $H_1$ term, while the remaining part is included in the $H_2$ term, thus giving \begin{equation} H_2=V \sum_{<ij>\sigma} a_{i\sigma}^{\dagger} a_{j\sigma} (a_{i\bar{\sigma}}^{\dagger} a_{i\bar{\sigma}}-\frac{\nu}{2})(a_{j\bar{\sigma}}^{\dagger}a_{j\bar{\sigma}} -\frac{\nu}{2}) \end{equation} with $\nu=<n_i>$ being the average occupation of the site. \section{The Cooper instability in the occupation-dependent hopping Hamiltonians} The Cooper instability realizes at certain temperature $T=T_c$ as a singularity in a two-particle scattering amplitude at zero total momentum. Let's introduce a function \begin{equation} \Gamma(p_1 p_2,\tau - \tau^{ \prime})=<T_{\tau} a_{{\bf p}_1 \uparrow} (\tau) a_{-{\bf p}_1 \downarrow}(\tau) \bar{a}_{-{\bf p}_2 \downarrow}(\tau^{ \prime}) \bar{a}_{{\bf p}_2 \uparrow} (\tau^{ \prime})> \end{equation} where $\bar{a}_{{\bf p}\alpha}(\tau) = \exp(H\tau) a_{{\bf p} \alpha}^{\dagger} \exp(-H\tau)$, $a_{{\bf p}\alpha} =\exp(H\tau) a_{{\bf p}\alpha} \exp(-H\tau)$ are the imaginary time $(\tau)$ creation and annihilation operators. At ${\bf p}_1=-{\bf p}_2$, ${\bf p}_3=-{\bf p}_4$, the kernel of $\Gamma_{\alpha \beta \gamma \delta}$ is proportional to $G_{\alpha \beta}^x$ $G_{\gamma \delta}^y$ ($G$ is one-electron Green function). We keep notation $\Gamma({\bf p p}^{\prime})$ for such a reduced Green function specifying only momenta ${\bf p}={\bf p}_1=-{\bf p}_2$ and ${\bf p}^{\prime}={\bf p}_3=-{\bf p}_4$. By assuming temporarily $V=0$, this Hamiltonian results in an equation for the Fourier transform $\Gamma(p,p^{\prime},\Omega)$ \begin{equation} \Gamma({\bf p},{\bf p}^{\prime}, \Omega)= \Gamma^0({\bf p}, {\bf p}^{\prime})- T \sum_{\omega} \sum_{{\bf{k}}} \Gamma^0({\bf{p}}, {\bf{k}})G_{\omega}({\bf{k}})G_{-\omega+\Omega}(-{\bf{k}}) \Gamma({\bf{k}},{\bf{p}^{\prime}},\Omega) \end{equation} corresponding to summation of Feynmann graphs shown in Figure 2. In the above formulas, $\omega=(2n+1)\pi T$ and $\Omega=2\pi m T$ ($n,m$ integers) are the discrete odd and even frequencies of the thermodynamic perturbation theory \cite{abrikosov1}. $G({\bf k}, \omega)$ is a one-particle Green function in a Fourier representation \begin{equation} G({\bf k}, \omega)= \frac{1}{\xi_k-i\omega}. \end{equation} Diagrams of Figure 2 are singular since equal momenta of two parallel running lines bring together singularities of both Green functions $G({\bf k},\omega)$ and $G(-{\bf k}, \omega)$. 6-vertex interaction, Eq.(8), is not generally considered in the theories of strongly-correlated fermionic systems. Such interaction also results in singular diagrams for ${\bf p} \rightarrow -{\bf p}$ scattering shown in Figure 3. Since a closed loop in this figure does not carry any momentum to the vertex, it reduces to the average value of $\bar G$ which in turn is the average of the number operator, $<a^{\dagger} a>$. Taking into consideration of such diagrams is equivalent to replacing one of the $n_i$'s in Eq.(8) to its thermodynamical average $\nu=<a_{i \sigma}^{\dagger}a_{i \sigma}>$. Then the $V$ term can be added to the renormalized value of $W$, \begin{eqnarray*} W \rightarrow W+ \frac{1}{2} \nu V \end{eqnarray*} We will check by numeric analysis in Sec. III to which extent such an approximation may be justified. Solution to Eq.(16) can be received by putting \begin{equation} \Gamma({\bf p},{\bf p}^{\prime}, \Omega)=A(\Omega)+B_1(\Omega)\sigma_{\bf p} +B_2(\Omega)\sigma_{{\bf p}^{\prime}}+C(\Omega)\sigma_{\bf p} \sigma_{{\bf p}^{\prime}}. \end{equation} Substituting this expression into Eq.(16) and introducing the quantities \begin{equation} S_n(\Omega)=T\sum_{\omega}\sum_{\bf k} \sigma_{\bf k}^{n}G_{\omega}({\bf k}) G_{-\omega+\Omega}(-{\bf k}) \end{equation} we receive a system of coupled equations for $A,B_1,B_2,C$ \begin{eqnarray} \left( \begin{array}{cccc} 1+U S_0 + \tilde{W} S_1 & US_1+{\tilde W}S_2 & 0 & 0 \\ {\tilde W}S_0 & 1+{\tilde W} S_1 & 0 & 0 \\ 0 & 0 & 1+U S_0 +{\tilde W}S_1 & U S_1+{\tilde W} S_2 \\ 0 & 0 & {\tilde W} S_0 & 1+{\tilde W} S_1 \end{array} \right) & & \left( \begin{array}{c} A \\ B_1 \\ B_2 \\ C \end{array} \right)= \left( \begin{array}{c} U \\ \tilde{W} \\ \tilde{W} \\ 0 \end{array} \right) \end{eqnarray} where $\tilde{W}=W+\frac{1}{2} \nu V$, which are solved to give \begin{equation} A=\frac{U-\tilde{W}^2S_2}{D} , \; B_1=B_2=\frac{\tilde{W}(1+\tilde{W}S_1)}{D} , \; C=-\frac{\tilde{W}^2S_0}{D} \end{equation} where $D$ is a determinant \begin{equation} D=\left| \begin{array}{cc} 1+US_0+\tilde{W}S_1 & US_1+\tilde{W}S_2 \\ WS_0 & 1+WS_1 \end{array} \right| . \end{equation} The determinant becomes zero at some temperature which means an instability in the two-particle scattering amplitude ($\Gamma\rightarrow\infty$). This temperature is the superconducting transition temperature $T_c$. At $T_c$, Eq.(16) is singular, which means that two-particle scattering amplitude gets infinite. Below $T_c$, the finite value of $\Gamma$ is established by including the non-zero thermal averages (the order parameters), $<a_{\bf p}^{\dagger} a_{-\bf p}^{\dagger}>$, $<a_{\bf p} a_{-\bf p}>$. We first analyze the case of non-retarded, non-contraction interaction $U$, and after that will consider the effect of the occupation-dependent hopping terms, $V$ and $W$. \subsection{Direct non-retarded interaction} Neglecting contraction parameters $V,W$, solution to Eq.(16) reduces to \begin{equation} -\frac{1}{U}=T\sum_{\omega}\sum_{\bf k} \frac{1}{\xi_{\bf k}^2+\omega^2} \end{equation} which after the summation over the discrete frequencies reduces to the conventional BCS equation (at negative $U$) \begin{equation} \frac{1}{|U|}=\sum_{\bf k} \frac{1-2n_{\bf k}}{2\xi_{\bf k}}, \end{equation} with $n_{\bf k}=(\exp(\beta\xi_{\bf k})+1)^{-1}$. At finite frequency $\Omega$, Eq.(23) reduces to \begin{equation} \ln \frac{T}{T_c}=T \sum_{\omega}\int_{-E_1}^{E_2} d\xi \frac{-i \Omega} {(\xi^2+\omega^2)(\xi+i\omega+i\Omega)} \end{equation} where we replaced for simplicity an integration over the Brillouin, zone $\int d^3k$, by the integration over the energy assuming that the density of states near the Fermi energy $\mu$ is flat. $-E_1$ and $E_2$ are the lower and upper limits of integration equal to $-4 t-\mu$ and $4 t-\mu$, respectively. Such an approximation is not very bad since most singular contribution to integral comes from the point $\xi_p=0$ where the integrand is the largest. Above $T_c$, Eq.(25) determines the frequency of the order parameter relaxation \cite{abrahams1,gorkov1,kulik4}. There is a small change in this frequency compared to the BCS model in which limits of the integration $(-E_1,E_2)$ are symmetric with respect to the Fermi energy, and small in comparison to $\varepsilon_F$, therefore we briefly discuss it now. To receive a real-time relaxation frequency, Eq.(25) needs to be analytically continued to a real frequency domain from the discrete imaginary frequencies $i\omega_n=(2n+1)\pi i T$ \cite{abrikosov1}. Using the identity \begin{equation} T \sum_{\omega}\frac{1}{(\omega+i \xi_1)(\omega+i \xi_2)...(\omega+i \xi_n)} =(-i)^n\sum_{i=1}^{n} \prod_{i\neq j} \frac{n(\xi_i)}{\xi_i-\xi_j} \end{equation} where $n(\xi)$ is a Fermi function $n(\xi)=(\exp(\beta \xi)+1)^{-1}$ gives \begin{equation} \ln \frac{T}{T_c}= \frac{i \Omega}{2} \int_{-E_1}^{E_2} \frac{\tanh \frac{\xi}{2T}}{\xi(2\xi + i \Omega)} d\xi \end{equation} where \begin{equation} T_c=\frac{2 \gamma}{\pi} \sqrt{E_1E_2}\exp \left(- \frac{1}{N(\varepsilon_F)|U|} \right) , \: \ln \gamma=C=0.577. \end{equation} $C$ is Euler constant. Analytic continuation is now simple: we change $\Omega$ to $i(\omega-i \delta)$, $\delta=+0$, to receive a function which will be analytic in the upper half plane of complex $\omega$, $Im \omega>0$. The order parameter relaxation equation becomes \begin{equation} \left( \ln \frac{T}{T_c}-\frac{\omega}{4}\int_{-E_1}^{E_2}\frac{\tanh \frac {\xi}{2T}}{\xi(\xi-\frac{\omega}{2}+i\delta)}d\xi\right) \Delta=0. \end{equation} At $\omega \ll T_c$ and $T-T_c \ll T_c$, the real and imaginary parts of Eq.(29) are easily evaluated to give \begin{equation} \left( T-T_c-\frac{\pi i \omega}{8 T_c}+ \omega \frac{E_1-E_2}{4E_1E_2} \right) \Delta=0. \end{equation} Thus, the order parameter relaxation equation at $T > T_c$ becomes \begin{equation} (1+i \lambda) \frac{\partial\Delta}{\partial t} + \Gamma \Delta= 0 \end{equation} where \begin{equation} \Gamma = \frac{8}{\pi}(T-T_c), \:\: \lambda=\frac{2(E_1-E_2)}{\pi E_1E_2} T_c. \end{equation} In comparison to the BCS theory in which $E_1=E_2=\omega_D$ ($\omega_D$ is the Debye frequency) and therefore $\lambda=0$, we receive the relaxation which has a non-zero ``inductive'' component, $-i \lambda \Gamma$. Typically, $E_1 \sim E_2 \sim \varepsilon_F$ and therefore $|\lambda|$ is a small quantity. It increases however near the low ($\nu \ll 1$) or near the maximal ($\nu \simeq 2$) occupation where $E_1$ or $E_2$ become small. Such mode of relaxation is specific to a non-retarded (non-phonon) interaction which is not symmetric near $\varepsilon_F$ and spans over the large volume of the ${\bf k}$-space rather than is restricted to a narrow energy $\omega_D \ll \varepsilon_F$ near the Fermi energy. \subsection{Occupation-dependent hopping instability and relaxation} Neglecting direct interaction, we put $U=0$ in Eq.(22) and receive \begin{equation} -\frac{1}{\tilde{W}}=S_1(\omega) \pm \sqrt{S_0(\omega)S_2(\omega)} \end{equation} where at finite frequency $\omega$ \begin{equation} S_n(\omega)=N(\varepsilon_F)T \sum_{\omega} \int_{-E_1}^{E_2} \left(\frac{\xi+\mu}{-t}\right)^n \frac{\tanh \frac{\xi}{2T}}{2\xi-\omega +i \delta} d\xi . \end{equation} Putting $\omega=0$ we receive from Eq.(33) a transition temperature $T_c$. The equation has a solution at $\tilde{W} < 0$, $\mu < 0$, or at $\tilde{W}>0$, $\mu>0$ (we assume that $t>0$). Plus or minus sign is chosen to receive the maximal value of $T_c$ (the second solution corresponding to smaller $T$, then, has to be disregarded since at $T<T_c$ the order parameter will be finite and therefore Eqs.(20)-(22) do not apply). This gives an expression for $T_c$ \begin{equation} T_c=\frac{2\gamma}{\pi} \sqrt{E_1E_2} \exp \left[ \frac{E_1-E_2}{2|\mu|t} (t-|\mu|)+\frac{E_2^2-E_1^2}{8\mu^2}\right] \exp \left(-\frac{t}{2|\tilde{W}|N(\varepsilon_F)}\right) \end{equation} where $\mu<0$, $\tilde{W}<0$ (second exponent is dominating the first one in the weak coupling limit $\tilde{W} \rightarrow 0$). Real and imaginary parts of $S_n(\omega)$ are calculated at $\omega \ll T_c$ \begin{eqnarray} ImS_n(\omega) \simeq -\frac{\pi\omega}{8T_c}\left( -\frac{\mu}{t} \right)^n N(\varepsilon_F)\\ ReS_n(\omega)=\frac{\omega}{4}N(\varepsilon_F) \left(-\frac{\mu}{t} \right)^n \times \left\{ \begin{array}{ll} \frac{E_2-E_1}{E_1E_2}, & n=0\\ \frac{E_2-E_1}{E_1E_2}+ \frac{2}{\mu} \ln \frac{\gamma \sqrt{E_1E_2}}{T_c}, & n=1\\ \frac{E_2-E_1}{E_1E_2}+\frac{E_1-E_2}{\mu^2}+ \frac{2}{\mu} \ln \frac{\gamma \sqrt{E_1E_2}}{T_c}, & n=2 \end{array} \right. \end{eqnarray} Equation for $\lambda$ is received with a value larger than the previous one (Eq.(32)) \begin{equation} \lambda \simeq \frac{T_c}{\mu} \left( 3 \ln \frac{2 \gamma \sqrt{E_1E_2}}{\pi T_c}+ \frac{2 \mu(E_2-E_1)}{E_1E_2}+\frac{E_1-E_2}{2\mu} \right). \end{equation} Eigenvalue equation gives the ${\bf p}$-dependence of the two particle correlator $\Gamma({\bf p},{\bf p}^{\prime})= \\ <a_{{\bf p} \uparrow}^{\dagger} a_{-{\bf p} \downarrow}^{\dagger} a_{-{\bf p}^{\prime} \downarrow} a_{{\bf p}^{\prime} \uparrow} >$ near $T_c$ \begin{equation} \Gamma({\bf p}, {\bf p}^{\prime})=C \left[ S_2-S_1(\sigma_{\bf p} + \sigma_{{\bf p}^{\prime}}) + S_0 \sigma_{\bf p} \sigma_{{\bf p}^{\prime}} \right]. \end{equation} Since $C$ diverges at $T_c$, this determines that order parameter becomes macroscopic at $T<T_c$. Then, the pair creation operator, $a_{\bf p}^{\dagger} a_{-\bf p}^{\dagger}$, will almost be a number, i.e., we may decompose Eq.(39) into a product \begin{equation} \Delta_{\bf p}^{\star} \Delta_{\bf p} = <a_{{\bf p} \uparrow}^{\dagger} a_{-{\bf p} \downarrow}^{\dagger}><a_{-{\bf p}^{\prime} \downarrow} a_{{\bf p}^{\prime} \uparrow}> \end{equation} and, to be consistent with the ${\bf p}$, ${\bf p}^{\prime}$ dependences, by putting $\xi_{\bf p}=\xi_{{\bf p}^{\prime}}$ we receive \begin{equation} \Delta_{\bf p}=C_1 \left( \exp(i \theta/2) \sqrt{S_2(0)}+\exp(-i \theta /2) \sqrt{S_0(0)} \right) \exp(i\varphi) \end{equation} where \begin{equation} \cos \theta= - S_1(0)/\sqrt{S_0(0)S_2(0)} \end{equation} and $\varphi$ is an overall phase which is irrelevant for a single superconductor but is important for calculating currents in multiple or weakly coupled superconductors. Therefore, system undergoes a pairing transition at temperature found from the Eq.(35). Since the pairs are charged, the state below $T_c$ can not be non-superconducting. We have not calculated the Meissner response but in the following section we present numerical calculation of flux quantization which supports the above statement. \section{Exact diagonalization of the occupation-dependent hopping Hamiltonians in finite cluster} We calculate the ground state energy of a cubic system as shown in Figure 4. A magnetic flux $\Phi$ is produced by a solenoid passing through the cube. Corners of the cube are the lattice sites, which can be occupied by electrons. With the inclusion of the magnetic flux, model Hamiltonian, Eq. 6, becomes \begin{eqnarray} H=-t \sum_{<ij>\sigma} a_{i \sigma}^{\dagger} a_{j \sigma} \exp(i \alpha_{ij})+h.c. +U \sum_i n_{i \uparrow}n_{i \downarrow} + \nonumber \\ + \sum_{<ij>\sigma} a_{i \sigma}^{\dagger} a_{j \sigma} \left[ V n_{i \bar{\sigma}} n_{j \bar{\sigma}} + W (n_{i \bar{\sigma}}+n_{j \bar{\sigma}}) \right] \exp(i \alpha_{ij}) +h.c. \end{eqnarray} where \begin{equation} \alpha_{ij}=(2\pi/\Phi_0)\int_{{\bf r}_i}^{{\bf r}_j} {\bf A} \, d{\bf l} \end{equation} and $\Phi_0=hc/e$ is the magnetic flux quantum. Throughout the calculations we take $t=1$. We start with constructing the model Hamiltonian. In a Hilbert space of one electron \begin{eqnarray} a=\left( \begin{array}{lr} 0 & 1 \\ 0 & 0 \end{array} \right), \; a^{\dagger}=\left( \begin{array}{lr} 0 & 0 \\ 1 & 0 \end{array} \right) . \end{eqnarray} with a basis specified as $\psi_0=(0,1)$ for the ground state ($n=0$) and $\psi_1=(1,0)$ for the excited state ($n=1$). In case of $N$ states operator of annihilation $a_n$ takes the form \begin{equation} a_n=s^{n-1}\otimes a \otimes e^{N-n} \end{equation} where $e$ is the unit matrix and $s$ is unitary matrix \begin{eqnarray} e=\left( \begin{array}{lr} 1 & 0 \\ 0 & 1 \end{array} \right), \; s=\left( \begin{array}{lr} 1 & 0 \\ 0 & -1 \end{array} \right) \end{eqnarray} and $\otimes$ stands for the Kronecker matrix multiplication. Explicitly, we have \begin{eqnarray*} a_1&=&a \otimes e \otimes e \otimes e \ldots \otimes e \\ a_2&=&s \otimes a \otimes e \otimes e \ldots \otimes e \\ &\ldots& \\ a_N&=&s \otimes s \otimes s \ldots \otimes s \otimes a \end{eqnarray*} Thus, for example, for two states \begin{eqnarray} a_1=\left( \begin{array}{cccc} 0 & 1 & 0 & 0 \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 1 \\ 0 & 0 & 0 & 0 \end{array} \right), \; a_2=\left( \begin{array}{cccc} 0 & 0 & 1 & 0 \\ 0 & 0 & 0 &-1 \\ 0 & 0 & 0 & 0 \\ 0 & 0 & 0 & 0 \end{array} \right) . \end{eqnarray} These matrices, which are annihilation operators, and corresponding Hermitian conjugate matrices, which are the creation operators, satisfy the Fermi anti-commutation relation. These operators are sparse matrices with only $N/2$ non-zero elements, which are equal to $\pm 1$. Next we solve the Schr\"{o}dinger equation $H\psi=E\psi$. We implemented a novel algorithm for solving such sparse systems, which will be described elsewhere. The cubic cluster within the Hubbard Hamiltonian and no external flux applied to the system was studied previously by Callaway et. al. \cite{callaway1}. Quantum Monte Carlo methods applicable to large systems within the Hubbard model (both attractive and repulsive), but not the occupation-dependent hopping Hamiltonians, are reviewed in a paper of Dagotto \cite{dagotto1}. \subsection{The number parity effect} Superconductivity reveals itself in the lowering of the ground state energy as electrons get paired. Therefore the energy needs to be minimal for even number of electrons $n$ and will attain a larger value when $n$ is odd. We consider a ``gap" parameter \cite{matveev1} \begin{equation} \Delta_l=E_{2l+1}-\frac{1}{2}\left(E_{2l}+E_{2l+2}\right) \end{equation} as a possible ``signature" of superconductivity (where $E_m$ corresponds to the ground state energy for $m$ fermions). For all interaction parameters set to zero ($U=V=W=0$), no sign of pairing is observed. To check our analytic results of Sec. IIb and the argument following Eq.(34), we calculated $\Delta$ above and below the half-filling ($n=8$ in case of cubic cluster). Below the half-filling chemical potential is negative ($\mu<0$) and above the half-filling it is positive ($\mu>0$). We first checked that the $W\rightarrow 0^+$, $W\rightarrow 0^-$ and $V\rightarrow 0^+$, $V\rightarrow 0^-$ calculation is consistent with an exact solution available for a non-interacting system of $n$ electrons. We then test our program for the case of negative-$U$ Hubbard Hamiltonian ($U<0, \, V=0,\, W=0$) which is known to be superconducting (e.g. Refs. 22,23). Positive-$U$ Hubbard model does not show any sign of superconductivity, in disagreement with some statements in the literature \cite{pao1}. Our calculations can not disprove the (possible) non-pairing mechanisms of superconductivity but these seem to be unlikely models for the problem of superconductivity in oxides which clearly shows pairing of electrons (holes) in the Josephson effect and in the Abrikosov vortices. The relation $2eV=\hbar \omega$ is justified in the first case \cite{witt1} and flux quantum of a vortex is $hc/2e$ in the second \cite{gammel1}, both with the value of the charge equal to twice the electronic charge, $e$. Figure 5 shows the dependence of the ground state energy upon the number of particles in case of negative-$U$ and positive-$U$ Hubbard models assuming $V=0$ and $W=0$. Such dependences are typical for any value of $|U|$. There clearly is the pairing effect when $U<0$ and there is no sign of pairing at $U>0$. Tests for pairing in the contraction $V$, $W$-models ($V \neq 0, \, U=W=0$ and $W\neq 0,\,U=V=0$, respectively) are shown in Figs. 6,7. The results are in agreement with our perturbative calculation of Sec. II and with its extension for the intermediate and strong coupling limits $|V|\gtrsim t$, $|W|\gtrsim t$. Since chemical potential is negative below the half-filling and positive above the half-filling, there is no pairing in the former case ($\tilde{W} \rightarrow 0^+$) and there is a sign of pairing in the latter case ($\tilde{W} > 0$), in accord with the value of the effective coupling constant $\tilde{W} = W+\frac{1}{2}\nu V$. Similarly, for $\tilde{W} \rightarrow 0^-$ below the half-filling there is a sign of pairing ($\Delta \neq 0$) while above the half-filling there is no pairing. These results are summarized in Table 1. For larger values of the interaction parameters, the perturbative results do not remain applicable anymore. Figure 8b shows the dependence of the parity gap $\Delta$ on the strength of the interaction. From Figure 8 it is understood that the $W$ interaction introduces a ``signature" of pairing in a similar way as the negative-$U$ interaction does. The possibility of the ``contraction" pairing has been investigated formerly in the papers \cite{hirsch3,boyaci1}. \subsection{Flux quantization} Flux quantization is another signature of superconductivity which is a consequence of the Meissner effect. We also tested for the periodicity of the energy versus flux dependence with the period $\Phi_1=hc/2e$ as compared to the period $\Phi_0=hc/e$ in the non-interacting system \cite{kulik6,buttiker1}. Unfortunately, the even harmonics of $\Phi_0$-periodic dependence of the ground state energy (and related to it, the harmonics of the persistent current $J=-\partial E/ \partial \Phi$ \cite{kulik6,buttiker1}) may simulate the pairing in a non-superconductive system. Small-size (mesoscopic) system can mask the superconducting behavior \cite{dagotto1}. Flux quantization in Hubbard Hamiltonians was studied formerly in Refs. 29-31. We first demonstrate the behavior of the ground state energy with respect to flux, Figure 9. A characteristic feature of mesoscopic system suggests that addition of one extra particle to the system changes the sign of the derivative of the ground state energy with respect to magnetic flux at $\Phi=0$. That is, depending on the parity of the number of particles and on the number of sites, system can change from paramagnetic to diamagnetic state or vice versa. But this behavior is not always observed for the cubic geometry studied. Except the sign change from $n=2$ to $n=3$ and from $n=7$ to $n=8$, no such behavior is seen. As mentioned above, however, the $\Phi_1$-periodic component of the $E(\Phi)$ dependence begins to appear at the higher value of $n$ (Figure 9,c). For both contraction parameters equal to zero, i.e. $W=V=0$, we observe appearance of the $hc/2e$-periodic component for some values of $U$ (Figure 10). Even for positive (repulsive) values of $U$, it is possible to see a local minimum appearing at $\Phi=hc/2e$ (Figure 10b). This is in agreement with the authors' previous works \cite{boyaci1,ferretti1}. But this minimum, which does not lead to an exact periodicity of the ground state energy with a period $\Phi_0/2$, should not be attributed to superconductivity, this is rather a characteristic behavior in mesoscopic systems. For $U<0$ (while $W=V=0$), the expected mesoscopic behavior, that is the change of the sign of the slope of ground state energy at $\Phi=0$, starts to demonstrate itself (Figure 11). But this happens at sufficiently large absolute values of (negative) $U$. For other values of $U$, however, there is no such change. More pronounced $hc/2e$-periodic components are observed with the introduction of non-zero interaction parameters. The role of $W$ on the ground state energy, when both $U$ and $V$ are zero, is shown in Figure 12. Meanwhile setting both $U$ and $W$ to zero and observing the effect of the non-zero $V$ shows that $V$ does not play a role as significant as the other two interaction parameters do. There is not much difference in the behavior of the ground state energy upon magnetic flux between the zero and non-zero $V$ (for example $V=-1$) cases. \section{Conclusions} We studied the peculiarity of electron conduction in systems in which conduction band is derived from the atomic shells with a small number of electrons ($N_e$) in an atom. Such materials may include oxygen ($N_e=8$) in the oxides, carbon ($N_e=6$) in borocarbides (e.g. $LuNi_2B_2C$), hydrogen ($N_e=1$) in some metals (e.g., $Pd-H$). Some materials of this kind are superconductors. It was argued that the Coulomb effects within the atoms strongly influence the inter-atom wave function overlap between the atomic sites and therefore the electron hopping amplitude between the sites. The phenomenology of such conduction mechanism results in a novel, to the conventional solid state theory, Hamiltonians called the occupation-dependent-hopping (or contraction) Hamiltonians, specified with the two coupling parameters $V$, $W$. We then attempted a study of superconductivity in such systems within the BCS-type approach assuming the Cooper pairing of electrons. The weak-coupling limit allows determination of the range of parameters $V$, $W$ values and also of the in-site Coulomb interaction $U$ value which show the Cooper instability. The strong-coupling limit was addressed by a numeric calculation on finite clusters using the novel algorithm (of non-Lanczos type) for eigenvalues of large sparse matrices. One of the results of this numeric calculation was that the positive-$U$ Hubbard model, sometimes believed to be a candidate for high-$T_c$ superconductivity, does not comply with the goal. This work was partially supported by the Scientific and Technical Research Council of Turkey (T\"{U}B{\.I}TAK) through the BDP program.
\section{Introduction} GRB980519\ was detected on 1998 May 19, 12:20:13 UT by CGRO--BATSE and {\em Beppo}SAX--GRBM. It was in the field of view of the {\em Beppo}SAX--WFC 2, allowing an estimate of its position with a 3 arcmin error circle at R.A. = 23$^{\rm h}$22$^{\rm m}$15$^{\rm s}$ Dec. = $+77^\circ15\hbox{$\,.\!\!^{\prime}$} 0$ (J2000). The GRBM light curve (lasting $\simeq 30$ s) together with the spectral evolution in the 40--700 keV range is shown in Fig. \ref{grbmlc}. The soft--hard--soft evolution is evident in the plot. In the 2--27 keV WFC band a similar behaviour is observed but the light curve is much more structured than the high energy one and it lasts for $\simeq 180$ s; the emission starts $\simeq 50$ s before the GRB trigger and stops $\simeq 130$ s later (in 't Zand et al. 1998). \begin{figure}[ht] \centerline{ \psfig{figure=R44f1.eps,height=10cm} } \caption{ The 1 s GRBM light curve of GRB980519\ and its power-law photon spectral index evolution. } \label{grbmlc} \end{figure} The average spectrum of the GRBM data can be well fitted with a single power-law with a photon spectral index ($N_E\propto E^{-\alpha}$) $\alpha=1.62\pm 0.10$ with $\chi^2\simeq 1$. The 40--700 keV fluence (over 27 s) is $F_\gamma = (8.1\pm 0.5)\times 10^{-6}$ erg cm$^{-2}$ and the hardness ratio $f_{100-300/50-100} = 2.2 \pm 0.2$. Description of the method adopted for GRBM spectra deconvolution is reported by Amati et al. (this workshop). \section{{\em Beppo}SAX\ NFIs observation and discussion} A follow-up observation performed with the {\em Beppo}SAX\ NFIs started less than 10 hrs after the trigger and a weak, rapidly decaying X-ray source was detected (Nicastro et al. 1998a). The decay was not monotonic, but the low counting rate did not allow us to reconstruct a detailed light curve. Figure \ref{xdecay} shows the 2--10 keV (MECS) flux decay and three possible power-law decay fits $F_x \propto t^{-\delta}$. It can be seen that constraining the fit to connect with the last part of the WFC detected flux, gives a {\em minimum} decay slope $\delta = 1.67\pm 0.03$. On the other hand, if the first point of the NFIs observation is {\em not} a peak superimposed to the monotonic decay, then we have $\delta = 2.25\pm 0.22$. This can be considered a {\em maximum} decay slope. It is realistic to suppose that the {\em real} slope is close to $\delta = 1.83\pm 0.30$ obtained excluding the first NFIs point from the fit. In any case, this is the most rapid decay for all the 13 GRB afterglows detected so far, typical values ranging between $1.1 \div 1.4$. \begin{figure}[tb] \centerline{ \psfig{figure=R44f2.eps,height=10cm} } \caption{ 2--10 keV X-ray flux decay as seen by the {\em Beppo}SAX\ MECS. The upper limit to the flux during the burst and the last WFC measured values are shown. Three possible fitted power-law decays are reported. } \label{xdecay} \end{figure} Spectral fitting of the 0.1--10 keV LECS+MECS data using an absorbed power-law gave a quite soft photon index of $2.8^{+0.6}_{-0.5}$ and $N_H$ in the range $0.3-2\; \times 10^{22}$ cm$^{-2}$ (see Fig. \ref{toospe}). The 0.1--2 keV flux is $(7.9\pm 2.5)\times 10^{-14}$ erg cm$^{-2}$ s$^{-1}$ while the 2--10 keV flux is $(1.4\pm 0.3)\times 10^{-13}$ erg cm$^{-2}$ s$^{-1}$. Further details are given in Nicastro et al. (1998b). \begin{figure} \centerline{ \psfig{figure=R44f3.eps,width=8.7cm} } \caption{ LECS+MECS 0.1--10 keV spectrum of the GRB afterglow. It is $\alpha=2.8^{+0.6}_{-0.5}$ and $N_H\approx 0.3-2\; \times 10^{22}$ cm$^{-2}$. } \label{toospe} \end{figure} Optical observations, started as early as 8 hours after the GRB, resulted in the detection of the afterglow (OT) in $VR_cI_c$ bands. The power-law decays in all these bands were all consistent with $\delta \simeq 2.0$. Deep observations performed $\simeq 66$ days after the burst with the 6-m telescope of the SAO--RAS revealed that at the position of the OT there is a faint object, possibly a galaxy, of magnitude $R_c \simeq 26$ (Sokolov et al. 1998b). It is worth to note that for all 9 optically identified GRBs, there are indications of the presence of an underlying host galaxy (Hogg \& Fruchter 1998; Sokolov et al. 1998a). The afterglow was also detected in the radio band by the VLA (Frail et al. 1998) at R.A. = 23$^{\rm h}$22$^{\rm m}$$21\fs49$ Dec. = $+77^\circ15'43\hbox{$\,.\!\!^{\prime\prime}$} 2$ (J2000, $\pm 0\fs1$). \begin{acknowledgements} This research is supported by the Italian Space Agency (ASI) and Consiglio Nazionale delle Ricerche (CNR). BeppoSAX is a major program of ASI with participation of the Netherlands Agency for Aerospace Programs (NIVR). All authors warmly thank the extraordinary teams of the BeppoSAX Scientific Operation Center and Operation Control Center for their enthusiastic support to the GRB program. K. H. is grateful to the US SAX Guest Investigator program for support. \end{acknowledgements}
\section{Introduction} Quasiassociative algebras, originally discovered by Vinberg [8]--[10] and Koszul~[3] in the 1960's in the study of homogeneous convex cones, appear also as an underlying structure of those Lie algerbras that possess a phase space. Namely, for a given Lie algebra ${\cal G}$, the following three conditions are equivalent [5]: \begin{enumerate} \item[(i)] ${\cal G} = Lie ({\cal R})$ for some quasiassociative algebra ${\cal R}$; \item[(ii)] Let $\rho: {\cal G} \rightarrow \mbox{End}({\cal{G}}^*)$ be a representation, not necessarily coadjoint one, such that on the semidirect sum Lie algebra ${\cal G} {\mathop {\vartriangleright\!\! <}\limits_{\rho}} {\cal G}^*$, the symplectic form is a 2-cocycle; \item[(iii)] The natural Poisson bracket on the Lie algebra ${\cal G} {\mathop {\vartriangleright\!\! <}\limits_{\rho}} {\cal G}^*$ is compatible with the canoni\-cal Poisson bracket. \end{enumerate} Thus, the quasiassociative algebras form a natural category from the point of view of Classical and Quantum mechanics. A list of Lie algebras with a phase space, given in~[5], includes such non-evident cases as Lie algebras of vector f\/ields on ${\mathbf R}^n$ and current algebras. One of the principal Lie algebras of physical interest, the Virasoro algebra, has been, however, conspiciously under-privileged so far. Its underlying quasiassociative structure is treated in the next two Sections. (Still more Lie algebras with a phase space can be found in Chapter~2 in~[7].) Before leaving the phase-space perspective for more mathematical matters, let me make two comments. First, the category of quasiassociative algebras is closed with respect to the operation of phase-space extension, unlike the smaller category of associative algebras: if ${\cal R}$ is quasiassociative then so is T$^*{\cal R}$, where multiplication in T$^*{\cal R}$ is given by the formula~[5] \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm a}} \setcounter{equation}{0} \begin{equation} \left(\begin{array}{c} x \\ \bar x \end{array} \right) * \left(\begin{array}{c} y \\ \bar y \end{array} \right) = \left(\begin{array}{c} x * y \\ x * \bar y \end{array} \right), \qquad x, y \in {\cal R}, \quad \bar x, \bar y \in {\cal R}^* = \mbox{Hom} ({\cal R}, \ldots ), \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm b}} \setcounter{equation}{0} \begin{equation} \langle x * \bar y, y \rangle = - \langle \bar y, x * y \rangle . \end{equation} Second, if $\rho: {\cal{G}} \rightarrow \mbox{End} ({\cal G}^*)$ is the representation staring in the properties (ii) and (iii) above, then the associated quasiassociative multiplication on ${\cal G}$ is given by the formula \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \setcounter{equation}{1} \begin{equation} x * y = \rho^d (x) (y), \end{equation} where $\rho^d: {\cal G} \rightarrow \mbox{Eng} ({\cal G})$ is the representation dual to $\rho$. The condition for the symplectic form on ${\cal G} {\mathop {\vartriangleright\!\! <}\limits_{\rho}} {\cal G}^*$ to be a 2-cocycle is then equivalent to the property \begin{equation} \rho^d (x) (y) - \rho^d (y) (x) = [x, y], \qquad \forall \; x, y \in {\cal G}. \end{equation} Thus, \begin{equation} Lie ({\mathrm T}^* {\cal R}) = {\mathrm T}^* Lie ({\cal R}). \end{equation} The equation (1.3) appears also in a very dif\/ferent context, as the condition for the complex of dif\/ferential forms on the Universal Enveloping Algebra $U({\cal G})$ to be ghost-free (see [6], equations (7.4) and (7.5).) Turning back to the Virasoro algebra, we see from formula $(*)$ that we have what appears to be a central extension of the corresponding centerless quasiassociative multiplication \begin{equation} e_p * e_q = - {q (1 + \epsilon q) \over 1 + \epsilon (p + q)} e_{p + q}, \qquad p, q \in {\mathbf Z}, \end{equation} where $\epsilon$ can be treated as either a formal parameter or a number such that $\epsilon^{-1} \bar \in {\mathbf Z}$. The next Section contains a quick verif\/ication that formula (1.5) satisf\/ies the quasiassociativity property $(**)$. Section 3 is devoted to central extensions of quasiassociative algebras in general and the algebra (1.5) in particular, resulting in the formula $(*)$ from the Abstract. In Section~4 we re-interpret in the language of 2-cocycles the property of a bilinear form to provide a central extension of a quasiassociative algebra; this interpretation then leads to a complex on the space of cochains $C^n = \mbox{Hom} ({\cal R}^{\otimes n},\cdot)$. Section~5 generalizes this complex to the case $C^n = \mbox{Hom} ({\cal R}^{\otimes n}, {\cal M})$, where ${\cal R}$ acts nontrivally on ${\cal M}$. Section~6 deals with the dual objects, homology. The last Section~7 is devoted to dif\/ferential-variational versions of the preceding results, for the case when the centerless Virasoro algebra is replaced by the Lie algebra of vector f\/ields on the circle with the commutator \begin{equation} [X, Y] = X Y' - X'Y, \qquad {}' = d/d z, \end{equation} and the central extension is given by the Gelfand-Fuks 2-cocycle \begin{equation} \omega (X, Y) = \int X Y'''\; d z. \end{equation} Appendix~1 contains a short proof that the Virasoro algebra does not come from an associative one. Semi-direct sums of quasiassociative algebras are treated in Appendix~2. In Appendix~3 we prove that if $G$ is a connected Lie group whose Lie algebra ${\cal G}$ comes out of a quasiassociative algebra then the Lie algebra ${\cal D} (G)$ of vector f\/ields on $G$ also allows a quasiassociative representation. \setcounter{equation}{0} \section{The Centerless Virasoro Algebra} Suppose a space with a basis $ \{ e_p|p \in G$, a commutative ring$\}$ has the multiplication of the form \begin{equation} e_p * e_q = f(p, q) e_{p+q}. \end{equation} Then \begin{equation} \hspace*{-5pt}\begin{array}{l} e_p * (e_q * e_r) - (e_p * e_q) * e_r = f(q, r) e_p * e_{q+r} - f(p, q) e_{p + q} *e_r \vspace{2mm}\\ \qquad = [f(q,r) f(p,q+r) - f(p,q) f(p+q,r)] e_{p+q+r}, \end{array} \end{equation} so that the quasiassociativity condition $(**)$, the symmetry between $p$ and $q$, is equivalent to the relation \begin{equation} f (q, r) f(p, q+r) - f(p,q) f(p+q,r) = f (p, r) f(q, p+r) - f(q, p) f(p+q,r), \end{equation} which can be rewritten as \begin{equation} [f(p,q) - f(q, p)] f(p+q,r) = f(q, r) f(p, q+r) - f(p, r) f(q, p+r). \end{equation} By formula (1.5), \begin{equation} f(p,q) = - {q (1 + \epsilon q) \over 1 + \epsilon (p + q)}, \end{equation} and we have to check that this $f(p,q)$ satisf\/ies formula (2.4). First, \begin{equation} f(p, q) - f(q, p) = {1 \over 1 + \epsilon (p+q)} [- q (1 + \epsilon q) + p (1 + \epsilon p) ] = p - q, \end{equation} so that \begin{equation} e_p * e_q - e_q * e_p = (p - q) e_{p+q}, \end{equation} guaranteeing that the Lie algebra generated by formula (1.5) is indeed the centerless Virasoro algebra. Now, for the LHS of formula (2.4) we obtain \renewcommand{\theequation}{\arabic{section}.\arabic{equation}$\ell$} \setcounter{equation}{7} \begin{equation} (p-q) {-r (1 + \epsilon r) \over 1+\epsilon (p + q + r)}, \end{equation} while for the RHS of formula (2.4) we get \renewcommand{\theequation}{\arabic{section}.\arabic{equation}$r$} \setcounter{equation}{7} \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle {- r(1 + \epsilon r) \over 1 + \epsilon (q+r)} \cdot {-(q+r) [1 + \epsilon (q+r)] \over 1 + \epsilon (p + q + r)} - {-r (1 + \epsilon r) \over 1 + \epsilon (p + r)} \cdot {-(p+r) [1 + \epsilon (p + r)] \over 1 + \epsilon (p + q + r) } \vspace{3mm}\\ \displaystyle \qquad = {- r (1 + \epsilon r) \over 1 + \epsilon (p + q + r) } [ - (q + r) + (p + r)] = {- r(1 + \epsilon r) \over 1 + \epsilon (p + q+r) } (p - q), \end{array} \end{equation} and this is the same as formula (2.8$\ell$). \medskip \noindent {\bf Remark 2.9.} Formula (2.5) is not the only solution of the equation (2.4) satisfying the Lie boundary condition \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \setcounter{equation}{9} \begin{equation} f (p, q) - f (q, p) = p - q. \end{equation} For example, \begin{equation} f (p, q) = \lambda - q, \qquad \lambda = \mbox{const}, \end{equation} is also a solution. It does not alow a proper central extension, however. \setcounter{equation}{0} \section{Central Extensions of Quasiassociative Algebras} Let $K$ be a commutative ring over which our quasiassociative algebra ${\cal R}$ is an algebra. Let $\Omega: {\cal R} \times {\cal R} \rightarrow K$ be a bilinear form. It def\/ines a multiplication on the space $\tilde {\cal R} = {\cal R} \oplus K$, by the rule \begin{equation} \left(\begin{array}{c} a \\ \alpha \end{array} \right) * \left(\begin{array}{c} b \\ \beta \end{array} \right) = \left(\begin{array}{c} a * b \\ \Omega (a, b) \end{array}\right), \qquad a, b \in {\cal R}, \quad \alpha, \beta \in K. \end{equation} When is $\tilde {\cal R}$ quasiassociative? We have: \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \left(\begin{array}{c} a \\ \alpha \end{array} \right) * \left( \left( \begin{array}{c}b \\ \beta \end{array} \right) * \left( \begin{array}{c} c \\ \gamma \end{array} \right) \right) - \left( \left(\begin{array}{c} a \\ \alpha \end{array} \right) * \left(\begin{array}{c} b \\ \beta \end{array} \right) \right) * \left(\begin{array}{c} c \\ \gamma \end{array} \right) \vspace{3mm}\\ \displaystyle = \left(\begin{array}{c} a \\ \alpha \end{array} \right) * \left(\begin{array}{c} b * c \\ \Omega (b,c) \end{array} \right) - \left(\begin{array}{c} a * b \\ \Omega (a, b) \end{array}\right) * \left(\begin{array}{c} c \\ \gamma \end{array} \right) = \left(\begin{array}{c} a * (b * c) - (a * b) * c \\ \Omega(a, b * c) - \Omega (a * b, c) \end{array}\right). \end{array}\hspace{-6.73pt} \end{equation} Thus, $\tilde {\cal R}$ is quasiassociative if\/f \begin{equation} \Omega (a, b * c) - \Omega (a * b, c) = \Omega (b, a * c) - \Omega (b * a, c). \end{equation} This can be equivalently rewritten as \begin{equation} \Omega (b, a * c) - \Omega (a, b * c) + \Omega ([a, b], c) = 0, \end{equation} where $[a, b] = a * b - b* a$ is the commutator in the Lie algebra $Lie({\cal R})$. By construction, the bilinear form \begin{equation} \omega (a, b) = \Omega (a, b) - \Omega (b, a) \end{equation} def\/ines a central extension of the Lie algebra $Lie({\cal R}$); thus, $\omega$ is a 2-cocycle on this Lie algebra. While we are at it, let's look at {\it trivial} central extensions of ${\cal R}$. These are produced from the multiplication \begin{equation} \left(\begin{array}{c} a \\ \alpha \end{array} \right) {\mathop{*}\limits_t} \left(\begin{array}{c} b \\ \beta \end{array} \right) = \left(\begin{array}{c} a * b \\ 0 \end{array} \right) \end{equation} by linear transformations of the form \begin{equation} \Phi = \left(\begin{array}{cc} id & 0 \\ \langle u, \cdot \rangle & 1 \end{array} \right), \qquad u \in {\cal R}^*. \end{equation} Thus, trivial extensions look like \begin{equation} \hspace*{-5pt}\begin{array}{l} \left(\begin{array}{c} a \\ \alpha \end{array} \right) * \left(\begin{array}{c} b \\ \beta \end{array} \right) = \Phi \left( \Phi^{-1} \left( \begin{array}{c} a \\ \alpha \end{array} \right) {\mathop{*}\limits_t} \Phi^{-1} \left(\begin{array}{c} b \\ \beta \end{array} \right) \right) \vspace{3mm}\\ \displaystyle \qquad =\Phi \left( \left(\begin{array}{c} a \\ \cdots \end{array} \right) {\mathop{*}\limits_t} \left(\begin{array}{c}b \\ \cdots \end{array} \right) \right) = \Phi \left(\begin{array}{c} a * b \\ 0 \end{array} \right) = \left(\begin{array}{c} a * b \\ \langle u, a * b \rangle \end{array} \right), \end{array} \end{equation} so that trivial ``2-cocycles'' on ${\cal R}$ are of the form \begin{equation} \Omega (a, b) = \langle u, a * b\rangle , \qquad u \in {\cal R}^*. \end{equation} The award of the title ``cocycle'' to $\Omega$ will be justif\/ied in the next Section, where the criterion (3.4) is recast as \begin{equation} \delta \Omega (a, b, c) = 0. \end{equation} Similarly, central extensions dif\/fering by a trivial 2-cocycle are equivalent: if \[ \left(\begin{array}{c} a \\ \alpha \end{array} \right) {\mathop {*}\limits_1} \left(\begin{array}{c} b \\ \beta \end{array} \right) = \left(\begin{array}{c} a * b \\ \omega (a, b) + \langle u, a* b \rangle \end{array} \right) \] and \[ \left(\begin{array}{c} a \\ \alpha \end{array} \right) {\mathop{*}\limits_2} \left(\begin{array}{c} b \\ \beta \end{array} \right) = \left(\begin{array}{c} a * b \\ \omega (a, b) \end{array} \right) \] are two such extensions, then the transformation $\Phi$ (3.7) takes the second multiplication into the f\/irst one. Let us return to the case of the Virasoro algebra. Formula $(*)$ shows that we have a central extension \begin{equation} \Omega (e_p, e_q) = \varphi (p) \delta^0_{p +q}. \end{equation} \begin{equation} 2 \varphi (p) = p^3 - \epsilon^{-1} p^2 - p + \epsilon p^2 = - p (1 -\epsilon p) \left(1 + \epsilon^{-1} p\right). \end{equation} The condition (3.4), in the notation (2.1) and (3.11), becomes: \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle 0 = \Omega (e_q, e_p * e_r) - \Omega (e_p , e_q * e_r) + \Omega ([e_p, e_q], e_r) \vspace{2mm}\\ \displaystyle \qquad = f (p, r) \Omega (e_q, e_{p + r}) - f(q, r) \Omega (e_p, e_{q+r}) +(p-q) \Omega (e_{p+q}, e_r) \vspace{2mm}\\ \displaystyle \qquad = [f (p, r) \varphi (q) - f (q, r) \varphi (p) + (p - q) \varphi (p +q)] \delta_{p + q + r}^0, \end{array} \end{equation} which can be rewritten as \begin{equation} (p-q) \varphi (p+q) = \varphi (p) f(q, - p-q) - \varphi (q) f (p, - p-q). \end{equation} With $f(p,q)$ and $\varphi (p)$ given by formula (2.5) and (3.12) respectively, for the $2 \times LHS$ of formula (3.14) we get: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}$\ell$} \setcounter{equation}{14} \begin{equation} -(p-q) (p+q) [1 - \epsilon (p+q)] \left[1 + \epsilon^{-1} (p+q)\right], \end{equation} while for the $2 \times RHS$ of formula (3.14) we obtain: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}$r$} \setcounter{equation}{14} \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle - p (1 - \epsilon p) \left(1 + \epsilon^{-1} p\right) \frac{(p+q) [1-\epsilon (p+q)]}{1 - \epsilon p} \vspace{3mm}\\ \displaystyle \qquad + q (1 - \epsilon q) \left(1 + \epsilon^{-1} q\right) \frac{(p+q) [1 - \epsilon (p+q)]} {1 - \epsilon q} \vspace{3mm}\\ \displaystyle \qquad = (p+q) [1-\epsilon (p+q)] (q - p) \left[1 + \epsilon^{-1} (p+q)\right], \end{array} \end{equation} and this is the same as the expression $(3.15\ell)$. Thus, we get a central extension of the quasiassociative algebra (2.1), (2.5). It remains to check that the 2-cocycle $\omega $ (3.5) is indeed the one entering the Virasoro algebra. We have: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \setcounter{equation}{15} \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \omega (e_p, e_q) = \Omega (e_p, e_q) - \Omega (e_q, e_p) = [\varphi (p) - \varphi (q) ] \delta^0_{p+q} \vspace{2mm}\\ \displaystyle \qquad = {1 \over 2} \left\{ \left[p^3 - p + \left(\epsilon - \epsilon^{-1} \right) p^2 \right] - \left[(-p)^3 - (-p) + \left(\epsilon - \epsilon^{-1}\right) (-p)^2\right] \right\} \delta^0_{p+q} \vspace{2mm}\\ \displaystyle \qquad = \left(p^3-p\right) \delta^0_{p+q}. \end{array} \end{equation} \setcounter{equation}{0} \section{The Quasiassociative Complex} Let ${\cal M}$ be a $K$-module. Def\/ine the cochains on ${\cal R}$ with values of ${\cal M}$ as \begin{equation} C^n = C^n ({\cal R}, {\cal M}) = \mbox{Hom}_K ({\cal R}^{\otimes n} , {\cal M}), \qquad n \in {\mathbf N}; \quad C^0: = {\cal M}. \end{equation} In the preceding Section we in ef\/fect met two coboundary operators $\delta: C^n \rightarrow C^{n+1}$ for $n=1$ and $n=2$, in formulae (3.9) and (3.4) respectively: \begin{equation} \psi \in C^1 \quad \Rightarrow \quad \delta \psi (a_1, a_2) = \psi (a_1 * a_2), \end{equation} \begin{equation} \psi \in C^2 \quad \Rightarrow \quad \delta \psi (a_1, a_2, a_3) = \psi (a_2, a_1 * a_3) - \psi (a_1, a_2 * a_3) + \psi ([a_1, a_2], a_3). \end{equation} It is obvious that $\delta^2=0$ on $C^1$, and the roundabout way this equality was verif\/ied in the preceeding Section actually proves that \begin{equation} H^2({\cal R}) : = H^2 ({\cal R}, K) \end{equation} describes the $K$-module of isomorphism classes of 1-dimensional central extensions of ${\cal R}$ by $K$. Guided by formulas (4.2) and (4.3), we def\/ine the coboundary operator $\delta: C^n \rightarrow C^{n+1}$ for all $n \in {\mathbf Z}_+$, as follows \begin{equation} \delta= 0 \quad \mbox{on} \quad C^0; \end{equation} \[ \psi \in C^n, \quad n \geq 1 \quad \Rightarrow \] \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm a}} \setcounter{equation}{5} \begin{equation} \delta \psi (a_1,\ldots, a_n, a) = \sum^n_{i = 1} (-1)^{i+1} \psi (\ldots \hat i \ldots, a_i * a) + \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm b}} \setcounter{equation}{5} \begin{equation} \qquad + \sum_{1 \leq i < j \leq n} (-1)^{i+j+1} \psi ([a_i, a_j]\ldots \; \hat i \ldots \; \hat j \; \ldots \, a). \end{equation} The hat over the argument signif\/ies this argument's absence; the last sum (4.6b) is missing when $n<2$; the right-most argument, $a=a_{n+1}$, is considered on a dif\/ferent footing from the rest, $a_1, \ldots, a_n$. (We see that $H^1 ({\cal R}) = \{\mbox{Ann} ({\cal R} * {\cal R}) \subset {\cal R}^*\}$.) Before proceeding further, we need to make some minimal skewsymmetry observations. \medskip \noindent {\bf Def\/inition 4.7.} {\it Suppose $n \geq 3$. If $\kappa$ is such that $2 \leq \kappa < n$, then a cochain $\psi \in C^n$ is called $\kappa$-skewsymmetric if it is skewsymmetric in its first $\kappa$ arguments.} \medskip \noindent {\bf Proposition 4.8.} {\it (i) For $n=2$, $\delta (C^n)$ is 2-skewsymmetric;\\ (ii) Suppose $n \geq 3$ and $\psi \in C^n$; if $\psi$ is $\kappa$-skewsymmetric then so is $\delta \psi$.} \medskip \noindent {\bf Proof.} (i) Formula (4.3) makes the claim obvious for $n=2$;\\ (ii) For $n \geq 3$, the sums (4.6a) and (4.6b) each change sign under the transposition $(i, i+1)$ for all $i < \kappa $. \hfill $\blacksquare$ \medskip Thereafter we assume that all our cochains are $\kappa$-skewsymmetric for some f\/ixed $\kappa \geq 2$. \medskip \noindent {\bf Proposition 4.9.} {\it $\delta^2 = 0$ on 2-skewsymmetric cochains.} \medskip \noindent {\bf Proof.} Let $\psi \in C^n, n \geq 2$. Set $\nu = \delta \psi$: \[ \hspace*{-8.5pt} \nu (a_1,\ldots, a_n, z) = \sum^n_{i = 1} (-1)^{i+1} \psi(\ldots \hat i \ldots , iz) + \sum_{1 \leq i < j \leq n} (-1)^{i+j+1} \psi ([i,j] \ldots \hat i \ldots \hat j \ldots , z), \] where for brevity we write $iz$ insted of $a_i * z$, and $[i, j]$ instead of $[a_i, a_j]$. Then, \[ \delta \nu (y_1,\ldots, y_{n+1}, t) \] \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm a}} \setcounter{equation}{10} \begin{equation} \qquad \qquad = \sum^{n+1}_{s=1} (-1)^{s+1} \nu (\hat s, st) + \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm b}} \setcounter{equation}{10} \begin{equation} \qquad \qquad + \sum_{1 \leq p < q \leq n+1} (-1)^{p+q+1} \nu [(p, q] \hat p \hat q, t), \end{equation} where for further bref\/ity we now suppress the ``$\ldots$'' convention. We shall work out separately the expression (4.11a) and (4.11b). \noindent ({\it a}) \ We have: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \setcounter{equation}{11} \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \nu (\hat s,t) = \sum_{i<s} (-1)^{i+1} \psi (\hat i \hat s , i (st)) + \sum_{i > s} (-1)^i \psi (\hat s \hat i, i (st)) \vspace{3mm}\\ \displaystyle \qquad+ \sum_{i < j < s} (-1)^{i+j+1} \psi ([i, j] \hat i \hat j \hat s , st) \vspace{3mm}\\ \displaystyle \qquad + \sum_{i < s, j>s} (-1)^{i+j} \psi ([i, j] \hat i \hat s \hat j, st) + \sum_{s< i< j} (-1)^{i+j+1} \psi ([i, j] \hat s \hat i \hat j, st). \end{array} \end{equation} Multiplying all this by $(-1)^{s+1}$ and summing on $s$, we get \[ \{(4.11{\rm a})\} = \] \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm a}} \setcounter{equation}{12} \begin{equation} \qquad = \sum_{a < b} (-1)^{a+b}\{\psi (\hat a \hat b , a (bt)) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm b}} \setcounter{equation}{12} \begin{equation} \qquad \qquad - \psi (\hat a \hat b, b (at))\} \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm c}} \setcounter{equation}{12} \begin{equation} \qquad + \sum_{a < b< c} (-1)^{a+b+c} \{ \psi ([a, b] \hat a \hat b \hat c, ct) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm d}} \setcounter{equation}{12} \begin{equation} \qquad \qquad - \psi ([a,c] \hat a \hat b \hat c , bt) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm e}} \setcounter{equation}{12} \begin{equation} \qquad \qquad + \psi ([b,c] \hat a \hat b \hat c, at)\} ; \end{equation} \noindent ({\it b}) \ We have: \[ \hspace*{-5pt}\begin{array}{l} \displaystyle \nu ([p, q] \hat p \hat q , t) = \psi ( \hat p \hat q , [p, q] t) + \sum_{\alpha < p} (-1)^\alpha \psi ([p, q]) \hat \alpha \hat p \hat q, \alpha t) \vspace{3mm}\\ \displaystyle \qquad + \sum_{p < \alpha < q} (-1)^{\alpha + 1} \psi ([p, q] \hat p \hat \alpha \hat q , \alpha t) + \sum_{\alpha > q} (-1)^\alpha \psi ([p, q] \hat p \hat q \hat \alpha, \alpha t) \vspace{3mm}\\ \displaystyle \qquad + \sum_{\alpha < p} (-1)^{\alpha + 1} \psi ([[p, q], \alpha ] \hat \alpha \hat p \hat q , t) + \sum_{p <\alpha < q} (-1)^\alpha \psi ([[ p, q], \alpha] \hat p \hat \alpha \hat q , t) \end{array} \] \[ \hspace*{-5pt}\begin{array}{l} \displaystyle \qquad + \sum_{\alpha > q} (-1)^{\alpha + 1} \psi ([[ p, q], \alpha] \hat p \hat q \hat \alpha, t) + \sum_{i < j < p} (-1)^{i + j+1} \psi ([i, j], [p, q] \hat i \hat j \hat p \hat q , t) \vspace{3mm}\\ \displaystyle \qquad + \sum_{i< p, p< j< q} (-1)^{i+j} \psi [i, j], [p, q] \hat i \hat p \hat j \hat q , t) + \sum_{i<p, j>q} (-1)^{i + j+1} \psi ([i, j], [p, q] \hat i \hat p \hat q \hat j , t) \vspace{3mm}\\ \displaystyle \qquad + \sum_{p< i<j<q} (-1)^{i+j+1} \psi ([i, j], [p,q] \hat p \hat i \hat j \hat q , t) + \sum_{p<i< q<j} (-1)^{i+j} \psi ([i, j], [p, q] \hat p \hat i \hat q \hat j , t) \vspace{3mm}\\ \displaystyle \qquad+ \sum_{q< i<j} (-1)^{i+j+1} \psi ([i, j], [p, q] \hat p \hat q \hat i \hat j , t). \end{array} \] Multiplying this monstrocity by $(-1)^{p+q+1}$ and summing on $\{1 \leq p <q \leq n + 1\}$, we f\/ind: \[ \{(4.11{\rm b})\} = \] \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm a}} \setcounter{equation}{13} \begin{equation} \qquad = - \sum_{a < b} (-1)^{a+b} \psi ( \hat a \hat b, [a, b]t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm b}} \setcounter{equation}{13} \begin{equation} \qquad + \sum_{a < b < c} (-1)^{a + b+c} \{ - \psi ([b, c] \hat a \hat b \hat c , at) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm c}} \setcounter{equation}{13} \begin{equation} \qquad + \psi ([a, c] \hat a \hat b \hat c , bt) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm d}} \setcounter{equation}{13} \begin{equation} \qquad - \psi ([a, b] \hat a \hat b \hat c , ct) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm e}} \setcounter{equation}{13} \begin{equation} \qquad + \psi ([[b, c],a] \hat a \hat b \hat c , t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm f}} \setcounter{equation}{13} \begin{equation} \qquad - \psi ([[a, c],b] \hat a \hat b \hat c , t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm g}} \setcounter{equation}{13} \begin{equation} \qquad + \psi ([[a, b],c] \hat a \hat b \hat c , t)\} \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm h}} \setcounter{equation}{13} \begin{equation} \qquad + \sum_{a< b< c< d} (-1)^{a+b+c+d} \{\psi ([a, b], [c, d] \hat a \hat b \hat c \hat d , t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm i}} \setcounter{equation}{13} \begin{equation} \qquad - \psi ([a, c], [b, d] \hat a \hat b \hat c \hat d , t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm j}} \setcounter{equation}{13} \begin{equation} \qquad + \psi ([a, d], [b, c] \hat a \hat b \hat c \hat d,t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm k}} \setcounter{equation}{13} \begin{equation} \qquad + \psi ([b, c], [a, d] \hat a \hat b \hat c \hat d ,t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm l}} \setcounter{equation}{13} \begin{equation} \qquad - \psi ([b, d], [a, c] \hat a \hat b \hat c \hat d,t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm m}} \setcounter{equation}{13} \begin{equation} \qquad + \psi ([o, d] [a, b] \hat a \hat b \hat c \hat d, t)\}. \end{equation} Grouping various terms together, we organize the cancellation scheme as follows: \begin{enumerate} \item[1)] (4.13a,b) and (4.14a), because of the equality \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \setcounter{equation}{14} \begin{equation} a (bt) - b (at) = [a, b] t \end{equation} being the def\/ining relation $(**)$ of a quasiassociative algebra; \item[2)] (4.13c) and (4.14d); (4.13d) and (4.14c); (4.13e) and (4.14b); \item[3)] (4.14e,f,g) by virtue of the Jacobi identity; \item[4)] (4.14h,m); (4.14i,l); (4.14j,k); -- all by virtue of $\psi$ being 2-skewsymmetric. \hfill $\blacksquare$ \end{enumerate} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \setcounter{equation}{0} \section{The Quasiassociative Complex with Values in a Module} In this Section we generalize the coboundary operator $\delta: C^n\rightarrow C^{n+1}$ given by formula~(4.6), to the case where the quasiassociative algebra ${\cal R}$ acts nontrivially on ${\cal M}$, the space where cochains take values. Suppose $\chi: {\cal R} \rightarrow \mbox{End} ({\cal M})$ is a linear map. It is natural to call it a representation of ${\cal R}$ if it behaves the way the left multiplication in ${\cal R}$ does: \begin{equation} \chi (a) \chi (b)- \chi (a * b) = \chi (b) \chi(a) - \chi (b*a), \qquad \forall \; a, b \in {\cal R}. \end{equation} Since this can be rewritten as \begin{equation} [\chi (a), \chi(b)] = \chi ([a, b]), \end{equation} we simply have a representation of the underlying Lie algebra $Lie({\cal R})$. It is interesting that for the purpose of extending the chain complex (4.6) of the preceeding Section, this natural and proper def\/inition is insuf\/f\/icient; a stronger one is required. This insuf\/f\/iciency can be seen as follows. Let $\psi: {\cal R} \rightarrow {\cal M}$ be a 1-cochain. By formula (4.2), we should now have \begin{equation} \delta \psi (a, b) = \psi (a * b) + c_1 \chi (a) \psi (b) + c_2 \chi (b) \psi(a), \end{equation} with some constants $c_1$ and $c_2$. If we f\/ix $m \in C^0 = {\cal M}$ and consider the natural def\/inition for the operator $\delta: C^0 \rightarrow C^1$, \begin{equation} \delta m (a) = \chi (a) (m), \end{equation} then \begin{equation} \hspace*{-5pt}\begin{array}{l} \left(\delta^2 m\right) (a, b) = \delta m (a * b) + c_1 \chi (a) \delta m (b) + c_2 \chi (b) \delta m (a) \vspace{2mm}\\ \qquad = (\chi(a * b) + c_1 \chi (a) \chi (b) + c_2 \chi (b) \chi(a)) (m). \end{array} \end{equation} This expression has no reasons to vanish unless we change the def\/inition of (left) representation to read \begin{equation} \chi (a * b) = \chi (a) \chi (b), \qquad \forall \; a, b \in {\cal R}, \end{equation} and set $c_1 = -1$, $c_2 = 0$ in formula (5.3). (We can also adapt the dual point of view, def\/ining (right) representation by the condition \begin{equation} \chi (a * b) = - \chi (b) \chi (a), \end{equation} and setting $c_1 = 0$, $c_2 = 1$ in formula (5.3). But we won't pursue this avenue here, leaving it to the next Section.) Thus, \begin{equation} \delta \psi (a,b) = \psi (a * b) - a . \psi (b), \end{equation} where \begin{equation} a . (\cdot) : = \chi (a) (\cdot). \end{equation} All told, we def\/ine the coboundary operator $\delta: \ C^n \rightarrow C^{n+1} $ by the formula \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \delta \psi (a_1,\ldots, a_n, a) = \sum^n_{i = 1} (-1)^{i+1} [ \psi (\ldots \hat i \ldots, a_i * a) - a_i . \psi (\ldots\hat i \ldots, a)] \vspace{3mm}\\ \displaystyle \qquad + \sum_{1 \leq i < j \leq n} (-1)^{i+j+1} \psi ([a_i, a_j] \ldots \hat i \ldots \hat j \ldots, a). \end{array} \end{equation} For $n=0$, formula (5.10) is to be understood as \begin{equation} \delta \psi (a_1) = - a_1 . \psi, \qquad \psi \in {\cal M}. \end{equation} The new extra sum in formula (5.10) doesn't destroy the property of $\delta$ to preserve $\kappa$-skewsymmeetry. \medskip \noindent {\bf Proposition 5.12.} $\delta^2 = 0.$ \medskip \noindent {\bf Proof.} We have seen above that $\delta^2 = 0$ on $C^0$, and it's easy to verify that $\delta^2 = 0$ on $C^1$ and $C^2$. So let $n \geq 3$. \setcounter{equation}{12} Setting \begin{equation} \delta = \delta^{\mbox{\scriptsize old}} + \delta^{\mbox{\scriptsize new}}, \end{equation} where $\delta^{\mbox{\scriptsize old}}$ is given by formula (4.6), and $\delta^{\mbox{\scriptsize new}}$ is given by formula \begin{equation} \delta^{\mbox{\scriptsize new}} \psi (a_1,\ldots, a_n, a) = \sum^n_{i = 1} (-1)^i a_i.\psi (\ldots \hat i \ldots,a), \end{equation} we have for $\nu = \delta \psi$: \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \nu (a_1, \ldots, a_n, z) = \sum^n_{i = 1} (-1)^{i+1} \psi (\hat i, iz) \vspace{3mm}\\ \displaystyle \qquad +\sum_{i < j} (-1)^{i+j+1} \psi ([i, j ] \hat i \hat j , z) + \sum^n_{i = 1} (-1)^i a_i . \psi ( \hat i ,z), \end{array} \end{equation} \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \delta \nu(y_1, \ldots, y_{n+1}, t) = \sum^{n+1}_{s=1} (-1)^{s+1} \nu (\hat s, st) \vspace{3mm}\\ \displaystyle \qquad + \sum_{p<q} (-1)^{p+q+1} \nu([p, q] \hat p \hat q , t) + \sum^{n+1}_{\ell = 1} (-1)^\ell y_\ell . \nu (\hat \ell, t). \end{array} \end{equation} We shall work out separately each of the three sums in the expression (5.16); since we have already verif\/ied in the preceding Section that $(\delta^{\mbox{\scriptsize old}})^2 = 0$, we shall only keep track of the extra terms coming out of the operator $\delta^{\mbox{\scriptsize old}} \delta^{\mbox{\scriptsize new}} + \delta^{\mbox{\scriptsize new}} \delta^{\mbox{\scriptsize old}} + (\delta^{\mbox{\scriptsize new}} )^2$. \noindent ({\it a}) \ We have: \[ \nu ( \hat s, st) \doteq \sum_{i <s} (-1)^i y_i . \psi (\hat i\hat s, st) +\sum_{i > s} (-1)^{i+1} y_i . \psi ( \hat s\hat i , st). \] Multiplying this by $(-1)^{s+1}$ and summing on $s$, we get \begin{equation} \{(5.16{\rm a})\} \doteq \sum_{a<b} (-1)^{a+b} \left[- y_a . \psi (\hat a \hat b, b t) + y_b . \psi (\hat a \hat b , a t) \right]; \end{equation} ({\it b}) \ We have: \[ \hspace*{-5pt}\begin{array}{l} \displaystyle \nu ([p, q] \hat p \hat q, t) \doteq - [y_p, y_q] . \psi (\hat p\hat q, t) + \sum_{\alpha< p} (-1)^{\alpha+1} y_\alpha . \psi ([p, q] \hat \alpha \hat p \hat q , t) \vspace{3mm}\\ \displaystyle \qquad + \sum_{p< \alpha<q} (-1)^\alpha y_\alpha . \psi ([p, q] \hat p \hat \alpha \hat q, t) + \sum_{\alpha > q} y_\alpha . \psi ([p, q] \hat p \hat q \hat \alpha, t). \end{array} \] Multiplying this by $(-1)^{p+q+1}$ and summing on $\{p<q\}$, we f\/ind: \begin{equation} \{(5.16{\rm b})\} \doteq \sum_{p<q} (-1)^{p+q} [y_p, y_q] . \psi ( \hat p \hat q, t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm a}} \setcounter{equation}{18} \begin{equation} \qquad + \sum_{a<b<c} (-1)^{a+b+c} \{ y_a . \psi ([b,c] \hat a \hat b \hat c,t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm b}} \setcounter{equation}{18} \begin{equation} \qquad \qquad -y_b . \psi ([a, c] \hat a \hat b \hat c ,t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm c}} \setcounter{equation}{18} \begin{equation} \qquad \qquad + y_c . \psi ([a, b] \hat a \hat b \hat c , t)\}; \end{equation} ({\it c}) \ We have: \[ \hspace*{-5pt}\begin{array}{l} \displaystyle y_\ell . \nu (\hat \ell , t)= y_\ell . \Biggl\{ \sum_{i < \ell} (-1)^{i+1} \psi (\hat i \hat \ell, it) + \sum _{i>\ell} (-1)^i \psi (\hat \ell \hat i, it) \vspace{3mm}\\ \displaystyle \qquad + \sum_{i<j<\ell} (-1)^{i+j+1} \psi([i, j]\hat i \hat j \hat \ell, t) + \sum_{i < \ell < j} (-1)^{i+j} \psi ([i, j] \hat i \hat \ell\hat j , t) \vspace{3mm}\\ \displaystyle \qquad + \sum_{\ell < i < j} (-1)^{i+j+1} \psi ([i, j] \hat \ell \hat i \hat j , t) + \sum_{i<\ell} (-1)^i y_i . \psi ( \hat i \hat \ell , t) +\sum_{i > \ell} (-1)^{i+1} y_i . \psi ( \hat \ell \hat i , t)\Biggr\} . \end{array} \] Multiplying all this by $(-1)^\ell$ and summing on $\ell$, we obtain: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \setcounter{equation}{19} \begin{equation} \{(5.16{\rm c})\}= \sum_{a<b} (-1)^{a+b} \{ - y_b . \psi (\hat a\hat b, at) + y_a . \psi ( \hat a \hat b, bt ) \} \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm a}} \setcounter{equation}{20} \begin{equation} \qquad + \sum_{a<b<c} (-1)^{a+b+c} \{ -y_c . \psi ([a, b] \hat a \hat b \hat c, t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm b}} \setcounter{equation}{20} \begin{equation} \qquad \qquad +y_b . \psi ([a, c] \hat a \hat b \hat c , t) \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm c}} \setcounter{equation}{20} \begin{equation} \qquad \qquad -y_a . \psi ([b, c] \hat a \hat b \hat c , t) \} \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \setcounter{equation}{21} \begin{equation} \qquad + \sum_{a<b} (-1)^{a+b} \{ y_b . (y_a . \psi (\hat a\hat b, t)) - y_a . (y_b . \psi (\hat a\hat b, t))\}. \end{equation} The cancellation scheme is: \begin{enumerate} \item[1)] (5.17) and (5.20); \item[2)] (5.18) and (5.22), since the action $\chi$ of ${\cal R}$ on ${\cal M}$ is a representation: \begin{equation} \chi ([a, b]) = [ \chi (a), \chi (b)], \qquad \forall \; a, b \in {\cal R}; \end{equation} \item[3)] (5.19) and (5.21). \end{enumerate} (Notice that $(\delta^{\mbox{\scriptsize new}})^2 \not= 0$.) \hfill $\blacksquare$ \medskip \noindent {\bf Remark 5.24.} The coboundary operator $\delta$ (5.10) does {\it not} reduce to the one of the Hochschild complex [2] when ${\cal R}$ is an associative algebra, even though the cochain spaces are identical in both cases. \medskip \noindent {\bf Remark 5.25.} When ${\cal M} = {\cal R}$ and the natural def\/inition of representation is used, one arrives at a new complex by considering deformations of the quasiassociative algebra ${\cal R}$, exactly like the Hochschild complex on $C^\bullet({\cal R}, {\cal R})$ is arrived at in the associative case~[1]. This new complex is closely related to the Hochschild one, and it is still dif\/ferent from the one constructed above. \setcounter{equation}{0} \section{Dual Point of View, Homology} The extended complex (5.10) of the preceding Section was based on the notion of representation of a quasiassociative algebra ${\cal R}$ as a linear map $\chi: {\cal R} \rightarrow \mbox{End} ({\cal M})$ satisfying the condition \begin{equation} \chi (a * b) = \chi (a) \chi (b). \end{equation} There was a second version of representation, formula (5.7): \begin{equation} \chi (a * b) = - \chi (b) \chi (a); \end{equation} this choice was left unexamined. Let's examine it now. These two choices lead to two dif\/ferent formulae for the coboundary operator $\delta: C^1 \rightarrow C^2$, \begin{equation} \delta \psi (a, b) = \psi (a * b) - \chi (a) \psi (b), \end{equation} \begin{equation} \delta \psi (a, b) = \psi (a * b) + \chi (b) \psi (a). \end{equation} The f\/irst direction was pursued in the preceding Section. The second one, as is easy to discover by considering the hypothetical map $\delta: C^2 \rightarrow C^3$, leads nowhere. Why is it so? Let ${\cal M}^* = \mbox{Hom}_K ({\cal M}, K)$ be the dual space to ${\cal M}$. Since ${\cal R}$ acts on ${\cal M}$, it also acts on ${\cal M}^*$ in the dual way: \begin{equation} \langle \chi^d (a) (m^*), m\rangle = - \langle m^*, \chi (a) (m)\rangle. \end{equation} Hence, \[ \hspace*{-5pt}\begin{array}{l} \langle \chi^d (a * b) (m^*), m\rangle = - \langle m^*, \chi (a * b) (m) \rangle \vspace{2mm}\\ \displaystyle \qquad = - \langle m^*,\chi (a) \chi (b) (m) \rangle = - \langle \chi^d (b) \chi^d (a) (m^*), m\rangle, \end{array} \] so that \begin{equation} \chi^d (a * b) = - \chi^d (b) \chi^d (a). \end{equation} Thus, our second version of representation, (6.2), is in fact dual to the f\/irst one, (6.1). Therefore, this def\/inition is suited not for cohomology but for the dual object, homology. Def\/ining the $n$-chains as \begin{equation} C_0 = C_0({\cal R}, {\cal N}) = {\cal N}; \qquad C_n = C_n ({\cal R}, {\cal N}) = {\cal N} \otimes {\cal R}^{\otimes n}, \qquad n \in {\mathbf N}, \end{equation} where ${\cal N}$ is a ${\cal R}$-module on which ${\cal R}$ acts according to formula (6.2): \begin{equation} {\bar n}^\bullet (a * b) = - {(\bar n}^\bullet a)^\bullet b, \qquad \bar n \in {\cal N}, \quad a, b \in {\cal R}, \end{equation} in the suggestive notation of the right action, we def\/ine the dif\/ferential $\partial: C_n \rightarrow C_{n-1}$ by the rule: \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \partial (\bar n \otimes a_1 \otimes \ldots \otimes a_n) = \sum^{n-1}_{i=1} (-1)^{i+1} \bar n \otimes \ldots \hat i \ldots a_i * a_n \vspace{3mm}\\ \displaystyle \qquad + \sum_{1 \leq i <j < n} (-1)^{i+j+1} \bar n \otimes [a_i, a_j] \ldots \hat i \ldots \hat j \ldots + \sum^{n-1}_{i=1} (-1)^{i+1} (n^\bullet a_i) \otimes \ldots \hat i \ldots, \vspace{3mm}\\ \displaystyle \partial (C_0) = 0, \qquad \partial (\bar n \otimes a) = {\bar n}^\bullet a. \end{array} \end{equation} Since this formula satisf\/ies the duality relation \begin{equation} \langle \partial \Psi, \psi \rangle = \langle \Psi, \delta \psi\rangle \end{equation} for the case $\Psi \in C_n ({\cal R}, {\cal N}) \approx (C^n ({\cal R}, {\cal N}^*))^* $, $\psi\in C^{n-1} ({\cal R}, {\cal N}^*)$, we have $\partial^2 = 0$ as a matter of course; it is assumed that the chains considered are $\kappa$-skewsymmetric for some $\kappa \geq 2$, exactly like the cochains. \medskip \noindent {\bf Remark 6.12.} The Hochschild coboundary operator on $C^1 =\mbox{Hom} ({\cal R}, {\cal M}) $ acts by the rule: \setcounter{equation}{12} \begin{equation} \delta \psi (a_1, a_2) = a_1 . \psi (a_2) - \psi (a_1 a_2) + \psi (a_1)^\bullet a_2, \end{equation} where ${\cal R}$ is associative, ${\cal M}$ is an ${\cal R}$-bimodule, and the right action of ${\cal R}$ on ${\cal M}$ is an anti-action from the pont of view of our def\/inition~(6.2). We see that formulae~(6.3) and~(6.4) each contribute about half to the Hochschild formula~(6.13). There must be some underlying reason for such split. \setcounter{equation}{0} \section{Dif\/ferential Algebra Viewpoint} Suppose our basic ring $K$ is a dif\/ferential ring, with a derivation $\partial: K \rightarrow K$. Then the formula \begin{equation} [X, Y] = XY' - X' Y, \qquad (\cdot)' = \partial (\cdot), \quad X, Y \in K, \end{equation} makes $K$ into a Lie algebra ${\cal D}_1 = {\cal D}_1 (K)$, the Lie algebra of vector f\/ields. The bilinear form $\omega$ on $K \times K$, \begin{equation} \omega (X, Y) = XY''' \end{equation} is skewsymmetric: \begin{equation} \omega (X, Y) \sim - \omega (Y, X), \end{equation} and is a generalized 2-cocycle on ${\cal D}_1$: \begin{equation} \omega ([X, Y], Z) + \omega ([Y, Z], X) + \omega ([Z, X], Y) \sim 0, \end{equation} where $(\cdot) \sim 0$ means that $(\cdot) \in \mbox{Im}\; \partial$. When \begin{equation} K = k \left[x, x^{-1}\right] \end{equation} and \begin{equation} \partial = d/dx, \end{equation} $k$ being some number f\/ield or such, the Lie algebra ${\cal D}_1$ is isomorphic to the centerless Virasoro algebra under identif\/ication \begin{equation} e_n = x^{1-n} {d \over dx}, \qquad X = \sum_n X_n e_n. \end{equation} As far as the Virasoro 2-cocycle is concerned, let \begin{equation} \mbox{Res}: \; k \left[x, x^{-1}\right] \rightarrow k \end{equation} be the map isolating the $x^{-1}$-coef\/f\/icient, so that \[ \mbox{Res} \circ \partial = 0. \] Then \begin{equation} \hspace*{-5pt}\begin{array}{l} \mbox{Res} (\omega (e_n, e_m)) = Res \left(x^{1-n} ( 1 - m) (-m) (-1-m) x^{-2-m} \right) \vspace{2mm}\\ \qquad = \delta^0_{n+m} (n+1) n (n-1) = \left(n^3 - n\right) \delta^0_{n+m}. \end{array} \end{equation} Below we construct a quasiassociative structure on $K = k\left[x, x^{-1}\right]$ and the corresponding generalized 2-cocycle on it, so that formulae~(1.5) and~(3.11) are recovered as localizations. Let \begin{equation} {\cal O} = x {d \over dx} -1 \end{equation} and set \begin{equation} u * v = (1 - \epsilon {\cal O})^{-1} x^{-1} u (1 - \epsilon {\cal O}) {\cal O}(v), \end{equation} \begin{equation} \hat \Omega (u, v) = x^{-3} {\cal O}^2 (1 + \epsilon {\cal O}) (u) \cdot v. \end{equation} Since \begin{equation} {\cal O} \left(x^{1-q}\right) = - q x^{1 - q}, \end{equation} we get \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle x^{1 - p} * x^{1-q} = (1 - \epsilon {\cal O})^{-1} \left(x^{-1} x^{1-p} (1 +\epsilon q) (-q) x^{1 - q}\right) \vspace{3mm}\\ \displaystyle \qquad = - q(1 - \epsilon q) (1 - \epsilon {\cal O})^{-1} x^{1-p-q} = -{q(1 + \epsilon q) \over 1 + \epsilon (p+q)} x^{1 - p - q}. \end{array} \end{equation} This is formula (1.5). It implies that we have a correct quasiassociative multiplication on $k\left[x, x^{-1}\right]$, with \begin{equation} u*v - v*u = uv' - u' v. \end{equation} The 2-cocycle story is more interesting. Recall how the notion of the {\it generalized} 2-cocycle on a Lie algebra, equation~(7.4), appears: from the classif\/ication of {\it affine} Hamiltonian operators, with the linear part being attached to a Lie algebra, say ${\cal G}$, and the constant part being a {\it generalized} 2-cocycle on this Lie algebra~[4]. Aposteriori one can put all this into a variational complex ([4], p.~204) $\delta: \ \mbox{Dif\/f}\;(\wedge^n {\cal G}, K)_v \rightarrow \ \mbox{Dif\/f}\; (\wedge^{n+1} {\cal G}, K)_v$, where subscript ``$v$'' signif\/ies that dif\/ferential forms dif\/fering by $Im\; \partial$ are to be identif\/ied; the generalized 2-cocycle condition~(7.4) is then simply \begin{equation} \delta \omega (X, Y, Z) \sim 0. \end{equation} We shall now apply the same variational leap-forward to the complex $C^\bullet ({\cal R}, K)$ of Section~5, considering cochains modulo $\mbox{Im}\; \partial$. A generalized 2-cocycle $\hat \Omega$ then satisf\/ies the dif\/ferential version of the equality~(3.4): \begin{equation} \delta \hat \Omega (u, v, w) = \hat \Omega (v, u*w) - \hat \Omega (u, v*w) + \hat \Omega ([u, v], w) \sim 0. \end{equation} It is unclear to me at the moment exactly what question such a variational 2-cocycle answers to, and repeated appeals to noncommutative dif\/ferential geometry in the sense of Allan Connes haven't helped so far; nevertheless, we have \setcounter{equation}{19} \medskip \noindent {\bf Proposition 7.19.} {\it (i) Let $\hat \Omega$ be a generalized 2-cocycle on a differential quasiassociative algebra ${\cal R}$. Then \begin{equation} \hat \omega (u, v) = \hat \Omega (u, v) - \hat \Omega (v, u) \end{equation} is a generalized 2-cocycle on the Lie algebra $Lie ({\cal R})$;\\ (ii) The symplectic form on $T^*{\cal R}$ is a generalized 2-cocycle on $T^*{\cal R}$.} \medskip \noindent {\bf Proof.} (i) We have, \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \hat \omega ([u, v], w) = \hat \Omega ([u, v], w) - \hat \Omega (w, [u, v]) \vspace{3mm}\\ \displaystyle \qquad {\mathop {\sim}\limits^{\mbox{\scriptsize \rm [by (7.18)]}}} \ \ \hat \Omega (u, v w) - \hat \Omega (v, uw) - \hat \Omega (w, uv - vu). \end{array} \end{equation} Hence, \[ \hspace*{-5pt}\begin{array}{l} \hat \omega ([u, v], w) + c.p. \sim (\hat \Omega (u, vw) + c.p.) - (\hat \Omega (v, uw) + c.p.) - (\hat \Omega (w, uv - vu) + c.p.) \vspace{3mm}\\ \displaystyle \qquad = (\hat \Omega (w, uv) + c.p.) - (\hat \Omega (w, vu) + c.p) - (\hat \Omega (w, uv - vu) + c.p.) = 0; \end{array} \] (ii) By formula (1.1), $T^*{\cal R}$ has the multiplication \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm a}} \setcounter{equation}{21} \begin{equation} \left(\begin{array}{c}u \\ \bar u \end{array} \right) * \left(\begin{array}{c}v \\ \bar v \end{array} \right) = \left(\begin{array}{c} u * v \\ u * \bar v \end{array} \right), \qquad u, v \in {\cal R}, \quad \bar u, \bar v \in {\cal R}^*, \end{equation} where \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm b}} \setcounter{equation}{21} \begin{equation} \langle u * \bar v, w \rangle \sim - \langle \bar v, u * w \rangle. \end{equation} The symplectic form $\hat \Omega $ is \[ \hat \Omega \left( \left(\begin{array}{c} u \\ \bar u \end{array} \right), \left(\begin{array}{c} v \\ \bar v \end{array} \right) \right) = \langle \bar u, v\rangle - \langle \bar v, u \rangle. \] Hence, \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm a}} \setcounter{equation}{22} \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \hat \Omega \left( \left(\begin{array}{c} v \\ \bar v \end{array} \right), \left(\begin{array}{c} u \\ \bar u \end{array} \right) * \left(\begin{array}{c} w \\ \bar w \end{array}\right) \right) = \langle \bar v, u * w \rangle - \langle u * \bar w, v\rangle \vspace{3mm}\\ \displaystyle \qquad \sim \langle \bar v, u * w \rangle + \langle \bar w, u * v \rangle, \end{array} \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm b}} \setcounter{equation}{22} \begin{equation} - \hat \Omega \left( \left(\begin{array}{c} u \\ \bar u \end{array}\right), \left(\begin{array}{c} v \\ \bar v \end{array} \right)*\left(\begin{array}{c} w \\ \bar w \end{array} \right) \right) \sim - \langle \bar u, v * w \rangle - \langle \bar w, v * u \rangle, \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm c}} \setcounter{equation}{22} \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \hat \Omega \left( \left[ \left(\begin{array}{c} u \\ \bar u \end{array} \right), \left(\begin{array}{c} v \\ \bar v \end{array} \right) \right], \left(\begin{array}{c} w\\ \bar w \end{array} \right) \right) = \Omega \left( \left(\begin{array}{c} [u, v] \\ u * \bar v - v * \bar u \end{array} \right), \left(\begin{array}{c} w \\ \bar w \end{array} \right) \right) \vspace{3mm}\\ \displaystyle \qquad = \langle u * \bar v - v * \bar u , w \rangle - \langle \bar w, [ u, v] \rangle \vspace{3mm}\\ \displaystyle \qquad \sim \ -\langle \bar v, u * w \rangle + \langle \bar u, v * w \rangle - \langle \bar w, u * v - v * u\rangle. \end{array} \end{equation} Adding the expressions (7.23a-c) up, we get zero. \hfill $\blacksquare$ \medskip Let us now verify that $\hat \Omega$ given by formula (7.13) is indeed a generalized 2-cocycle. We have: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \setcounter{equation}{23} \begin{equation} \hspace*{-5pt}\begin{array}{ll} 1) & \hat \Omega (v, u w) = x^{-3} {\cal O}^2 (1 + \epsilon {\cal O}) (v) \cdot (1 - \epsilon {\cal O})^{-1} x^{-1} u (1 - \epsilon {\cal O}) {\cal O}(w) \vspace{2mm}\\ & \displaystyle \qquad \sim [1+ \epsilon ({\cal O} + 3)]^{-1} x^{-3} {\cal O}^2 (1 + \epsilon {\cal O}) (v) \cdot x^{-1} u (1 - \epsilon {\cal O}) {\cal O} (w) \vspace{3mm}\\ & \displaystyle \qquad \sim \left\{ - ({\cal O} + 3 ) [1 + \epsilon ({\cal O} + 3) ] x^{-1} u [1 + \epsilon ({\cal O}+ 3)]^{-1} x^{-3} {\cal O}^2 (1 + \epsilon {\cal O}) (v) \right\} \cdot w, \end{array} \hspace{-16.8pt} \end{equation} where we used the universal relation \begin{equation} (1) A(2) \sim A^\dagger (1) \cdot (2) \end{equation} for the adjoint operator, and the particular relation \begin{equation} {\cal O}^\dagger = - ({\cal O} + 3) \end{equation} for our operator $\displaystyle {\cal O} = x {d \over dx} -1$. Now, since \begin{equation} ({\cal O} + 3) x^{-3} = x^{-3} {\cal O}, \end{equation} formula (7.24) can be rewritten as \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm a}} \setcounter{equation}{27} \begin{equation} \left\{ -x^{-4} ({\cal O}-1) [1+ \epsilon ({\cal O} -1)] u {\cal O}^2 (v) \right\} \cdot w; \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm b}} \setcounter{equation}{27} \begin{equation} \hspace*{-5pt}\begin{array}{ll} 2) & - \hat \Omega (u, vw) \sim \left\{ x^{-4} ({\cal O} - 1) [1 + \epsilon ({\cal O} -1)] v {\cal O}^2 (u) \right\} \cdot w; \end{array} \end{equation} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}{\rm c}} \setcounter{equation}{27} \begin{equation} \hspace*{-5pt}\begin{array}{ll} 3) & \hat \Omega ([u, v], w ) = \left\{ x^{-3} {\cal O}^2 (1 + \epsilon {\cal O}) (uv' -u' v) \right\} \cdot w. \end{array} \end{equation} Adding up the expressions (7.28a-c), we arrive at the equivalent relation to be verif\/ied: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \setcounter{equation}{28} \begin{equation} x^{-3} {\cal O}^2 (1 + \epsilon {\cal O}) (uv' - u'v) = x^{-4} ({\cal O} -1)[1 + \epsilon ({\cal O} - 1)] \left[ v {\cal O}^2 (u) - u {\cal O}^2 (v)\right]. \end{equation} Since \begin{equation} x^{-1} ({\cal O} -1) = {\cal O} x^{-1}, \end{equation} equality (7.29) reduces to \begin{equation} {\cal O} (uv' - u' v) = x^{-1} \left[u {\cal O}^2 (v) - v {\cal O}^2 (u)\right]. \end{equation} Now, \begin{equation} {\cal O}^2 = x^2 {d^2 \over dx^2} - x {d \over dx} + 1, \end{equation} so that \[ \hspace*{-5pt}\begin{array}{l} \displaystyle x^{-1} \left[u {\cal O}^2 (v) - v {\cal O}^2 (w)\right] = u (x v'' - v') -v (xu''- u') \vspace{3mm}\\ \displaystyle \qquad = x (uv'' - u'' v) - (u v' - u' v) = \left(x {d \over dx} - 1\right) (uv' - u' v) = {\cal O} (uv' - u'v). \end{array} \] It remains to perform the last step: to calculate $\mbox{Res}\; \hat \Omega \left(x^{1-p}, x^{1-q}\right)$ and to compare the result with the formulae (3.11,12). We have: \begin{equation} \hat \Omega \left(x^{1-p} , x^{1-q}\right) = x^{-3} {\cal O}^2 (1 + \epsilon {\cal O}) \left(x^{1-p} \right) \cdot x^{1-q} \ \ {\mathop{=}\limits^{\mbox{\scriptsize \rm [by (7.14)]}}} \ \ (-p)^2 (1 - \epsilon p) x^{-1-p-q}, \end{equation} so that \begin{equation} \mbox{Res}\; \hat \Omega \left(x^{1-p}, x^{1-q}\right) = p^2 (1 - \epsilon p) \delta^0_{p+q} = - \epsilon \left(p^3 - \epsilon^{-1} p^2\right) \delta^0_{p+q}. \end{equation} We see that we have to multiply $\hat \Omega$ by $- {1 \over 2} \epsilon^{-1}$, and also to add to it the trivial 2-cocycle proportional to the $*$ product. From formula~(7.15) we f\/ind: \begin{equation} \mbox{Res} \; x^{-2} \left(x^{1-p} * x^{1-q} \right) = p (1 - \epsilon p) \delta^0_{p+q} . \end{equation} Thus, the correctly normalized generalized 2-cocycle has the form \begin{equation} \hat \Omega^{\mbox{\scriptsize new}} (u, v) = - {1 \over 2} \epsilon^{-1} x^{-3} {\cal O}^2 (1 + \epsilon {\cal O}) (u) \cdot v - {1 \over 2} x^{-2} (1 - \epsilon {\cal O})^{-1} x^{-1} u (1 - \epsilon {\cal O}){\cal O} (v). \end{equation} \setcounter{equation}{36} \noindent {\bf Remark 7.36.} Consider the Lie algebra ${\cal D}_n = {\cal D}_n(K)$ ``of vector f\/ields on ${\mathbf R}^n$'', with the commutator \begin{equation} [X,Y]^i = \sum^n_{s=1} (X^s Y^i,_s - Y^s X^i,_s) , \qquad X, Y \in K^n, \end{equation} where \begin{equation} (\cdot ),_s = \partial_s (\cdot) , \end{equation} and $\partial_1,\ldots, \partial_n: \ K \rightarrow K$ are $n$ commuting derivations. Localizing $K$ as $k [x_1,\ldots, x_n, x_1^{-1}$, $\ldots, x_n^{-1}]$ and taking as the basis of ${\cal D}_n (K)$ \begin{equation} e^i_\sigma = x^{1 _{i}-\sigma} \partial_i = x_1^{-\sigma_{1}} \ldots x_n^{- \sigma_n} x_i \partial_i, \qquad \sigma \in {\mathbf Z}^n, \quad \partial_i = \partial/\partial x_{i }, \end{equation} we f\/ind the $n$-dimensional analog of the centerless Virasoro algebra: \begin{equation} [e^i_\sigma, e^j_\nu] = (\delta_{ij} - \nu_i) e^j_{\sigma + \nu} - (\delta_{ij} - \sigma_j) e^i_{\sigma + \nu}. \end{equation} This Lie algebra does not seem to have a quasiassoactive representation of the form~(1.5) for $n > 1$, but it does have a quasiassociative representation generalizing formula~(2.11): \begin{equation} e^i_\sigma * e^j_\nu = (\lambda \delta_{ij} - \nu_i) e^i_{\sigma + \nu}, \qquad \lambda = \mbox{const}. \end{equation} \renewcommand{\theequation}{{\rm A}1.\arabic{equation}} \setcounter{equation}{0} \subsection*{Appendix 1. Virasoro Algebra Does Not Come from an Associative One} Suppose we have a ${\mathbf Z}$-graded multiplication on the basis $\{e_p \mid p \in G$, a commutative ring$\}$, of the form \begin{equation} e_i e_j = g (i, j) e_{i + j}, \end{equation} such that \begin{equation} e_i e_j - e_j e_i = (i - j) e_{i+j}, \qquad \forall \; i, j \in G, \end{equation} and \begin{equation} (e_i e_j) e_k = e_i (e_j e_\kappa), \qquad \forall \; i, j, \kappa \in G. \end{equation} Let us show that such representation is impossible. We f\/irst rewrite the boundary condition (A1.2) as \begin{equation} g (i, j) - g(j, i) = i-j. \end{equation} Next, rewrite the associativity condition (A1.3) as \begin{equation} g(i,j) g(i+j, \kappa) = g (j, \kappa) g (i, j+\kappa). \end{equation} Now, set $j = \kappa = 0$ in formula (A1.5): \begin{equation} g (i, 0) [g (i, 0) - g (0, 0)] = 0. \end{equation} Further, set $j = i = 0$ in formula (A1.5): \begin{equation} g (0, \kappa) [g (0, \kappa) - g (0, 0)] = 0. \end{equation} Assume that $G$ has no zero divisors. From formula~(A1.6) we f\/ind: \begin{equation} g (i, 0) = 0 \quad \mbox{ or} \quad g(0,0), \end{equation} while formula (A1.7) yields: \begin{equation} g(0, \kappa) = 0 \quad \mbox{ or} \quad g(0, 0). \end{equation} The last two equations contradict the boundary condition~(A1.4): \[ g (r, 0) - g(0, r) = r. \] \renewcommand{\theequation}{{\rm A}2.\arabic{equation}} \setcounter{equation}{0} \subsection*{Appendix 2. Semidirect Sums of Quasiassociative Algebras} Let ${\cal R}$ and ${\cal U}$ be quasiassociative algebras, ${\cal G} = Lie ({\cal R})$, ${\cal H} = Lie ({\cal U})$. Let $\chi : {\cal G} \rightarrow \mbox{Der} ({\cal H})$ be a representation of ${\cal G}$. The semidirect sum Lie algebra ${\cal G} {\mathop {\vartriangleright\!\! <}\limits_{\chi}} {\cal H}$ is the vector space ${\cal G} \oplus {\cal H}$ with the commutator \begin{equation} \left[ \left(\begin{array}{c} a \\ u \end{array} \right), \left(\begin{array}{c} b \\ v \end{array}\right) \right] = \left(\begin{array}{c} [a, b] \\ a . v - b . u + [ u, v] \end{array} \right) , \qquad a, b \in {\cal G}, \quad u, v \in {\cal H}. \end{equation} Does the Lie algebra ${\cal G} {\mathop {\vartriangleright\!\! <}\limits_{\chi}} {\cal H}$ have a quasiassociative representation? \medskip \setcounter{equation}{2} \noindent {\bf Proposition A2.2.} {\it Let $\chi: \ Lie({\cal R}) \rightarrow \mbox{\rm Der} ({\cal U})$ be a representation. Define the semidirect sum ${\cal R} {\mathop {\vartriangleright\!\! <}\limits_{\chi}} {\cal U}$ as the space ${\cal R} \oplus {\cal U}$ with the multiplication \begin{equation} \left(\begin{array}{c} a \\ u \end{array} \right) * \left(\begin{array}{c} b \\ b \end{array} \right) = \left(\begin{array}{c} a * b \\ a . v + u * v \end{array} \right), \qquad a, b \in {\cal R}, \quad u, v \in {\cal U}. \end{equation} Then this multiplication is quasiassociative.} \medskip \noindent {\bf Proof.} Dropping the $*$ notation for brevity, we have \[ \hspace*{-14.2pt} \left(\begin{array}{c}a \\ u \end{array} \right) \left( \left(\begin{array}{c} b \\ v \end{array} \right) \left(\begin{array}{c} c \\ w \end{array}\right) \right) = \left(\begin{array}{c} a \\ u \end{array}\right) \left(\begin{array}{c} bc \\ b . w + v w \end{array} \right) = \left(\begin{array}{c} a (bc) \\ a . (b . w + v w ) + u (b . w + vw) \end{array} \right), \] \[ \left( \left(\begin{array}{c}a \\ u \end{array} \right) \left(\begin{array}{c} b \\ v \end{array}\right) \right) \left(\begin{array}{c}c \\ w \end{array} \right) = \left(\begin{array}{c} ab \\ a . v + u v \end{array} \right) \left(\begin{array}{c} c \\ w \end{array} \right) = \left(\begin{array}{c} (a b) c \\ (a b) . w + (a . v + uv )w \end{array} \right). \] Thus, we need to verify that \[ \hspace*{-5pt}\begin{array}{l} a . (b . w + vw) + u (b . w + vw) - (a b) . w - (a . v + uv) w \vspace{2mm}\\ \qquad = b . (a . w + uw) + v (a . w + uw) - (b a) . w - (b . u + vu) w. \end{array} \] This can be rewritten as $0 {\mathop{=}\limits^{?}}$ \renewcommand{\theequation}{{\rm A}2.\arabic{equation}{\rm a}} \setcounter{equation}{3} \begin{equation} a . (b . w) - b . (a . w) - ((a b) . w - (b a). w) \end{equation} \renewcommand{\theequation}{{\rm A}2.\arabic{equation}{\rm b}} \setcounter{equation}{3} \begin{equation} \qquad +a . (vw) - (a . v) w - v (a . w) \end{equation} \renewcommand{\theequation}{{\rm A}2.\arabic{equation}{\rm c}} \setcounter{equation}{3} \begin{equation} \qquad + u (b. w) + (b. u) w - b. (u w) \end{equation} \renewcommand{\theequation}{{\rm A}2.\arabic{equation}{\rm d}} \setcounter{equation}{3} \begin{equation} \qquad + u (vw) - (uv) w - v (u w) + (vu) w. \end{equation} The f\/irst sum vanishes since $\chi$ is a representation of $Lie \; ({\cal R})$; the second and third sums vanish since $Im (\chi) \subset \mbox{Der} ({\cal U})$; the fourth sum vanishes since ${\cal U}$ is quasiassociative. \hfill $\blacksquare$ \medskip \noindent {\bf Corollary A2.5.} {\it If ${\cal U}$ is abelian and $\chi: \ Lie ({\cal R}) \rightarrow \mbox{\rm End}({\cal U})$ is a representation, then ${\cal R} {\mathop{\vartriangleright\!\! <}\limits_{\chi}} {\cal U}$ is quasiassociative.} \medskip \noindent {\bf Proof.} $\mbox{Der} ({\cal U}) = \mbox{End} ({\cal U})$ for an abelian ${\cal U}$. \hfill $\blacksquare$ \medskip \noindent {\bf Example A2.6.} Consider the Ehrenfest Lie algebra ${\cal G} (A)$, where $A$ is an arbitrary matrix, and the commutators between basis elements are ([6], p.~274): \renewcommand{\theequation}{{\rm A}2.\arabic{equation}} \setcounter{equation}{6} \begin{equation} [e_i, e_j] = [\bar e_i, \bar e_j] = 0, \qquad [e_i, \bar e_j] =A_{ji}\bar e_j. \end{equation} In this case both ${\cal R}$ and ${\cal U}$ are vector spaces with trivial multiplication, \begin{equation} a * b = 0, \qquad u * v = 0, \qquad \forall \; a, b \in {\cal R}, \quad u, v \in {\cal U}, \end{equation} and the representation $\chi$ acts by the fule \begin{equation} e_i . \bar e_j = A_{ji} \bar e_j, \end{equation} It {\it is} a representation of the abelian Lie algebra $Lie({\cal R})$, since \begin{equation} e_i . (e_j . \bar e_\kappa) = e_j . (e_i . \bar e_\kappa) = A_{\kappa i}A_{\kappa j} \bar e_\kappa. \end{equation} Hence, the Ehrenfest Lie algebra ${\cal G} (A)$ (A2.7) comes out of the following quasiassociative multiplication: \begin{equation} e_i e_j = \bar e_i \bar e_j = \bar e_i e_j = 0, \qquad e_i \bar e_j = A_{ji} \bar e_j. \end{equation} \setcounter{equation}{12} \noindent{\bf Remark A2.12.} Proposition A2.2 shows that \begin{equation} Lie (R) {\mathop{\vartriangleright\!\! <}\limits_{\chi}} Lie ({\cal U}) = Lie ({\cal R} {\mathop{\vartriangleright\!\! <}\limits_{\chi}} {\cal U}) \end{equation} when ${\cal U}$ is abelian. Otherwise formula (A2.13) is not necessarily true since $\mbox{Der} ({\cal U})$ is, in general, smaller than $\mbox{Der} (Lie ({\cal U})):$ \medskip \noindent {\bf Proposition A2.14.} {\it (i) $\mbox{\rm Der} ({\cal U}) \subset \mbox{\rm Der} (Lie ({\cal U}))$;\\ (ii) If $\mbox{\rm Int}({\cal U}) \subset \mbox{\rm Der} ({\cal U})$ then ${\cal U}$ is associative. (Here $\mbox{\rm Int}({\cal U})$ denotes the space of maps $\{\mbox{\rm ad}_u: \ {\cal U} \rightarrow {\cal U} \ |u \in {\cal U} \}$.)} \medskip \noindent {\bf Proof.} (i) is well-known to be true for any algebra, not necessarily associative or quasiassociative one; (ii) $\mbox{\rm ad}_u$ is a derivation of $Lie({\cal U})$ no matter whether ${\cal U}$ is quasiassociative or not. For $\mbox{ad}_u$ to be a derivation of ${\cal U}$, we must have, for any $u, v, w \in {\cal U}$: \[ 0 = ad_u (vw) - (ad_u (v)) w -v ad_u (w) = u (vw) - (vw) u - (uv - vu) w - v (uw - wu) \] \renewcommand{\theequation}{{\rm A}2.\arabic{equation}{\rm a}} \setcounter{equation}{14} \begin{equation} \qquad = u (v w) - (uv) w - v (uw) + (vu) w \end{equation} \renewcommand{\theequation}{{\rm A}2.\arabic{equation}{\rm b}} \setcounter{equation}{14} \begin{equation} \qquad - (vw) u+ v (wu). \end{equation} The f\/irst sum vanishes since ${\cal U}$ is quasiassociative. The second sum vanishes if\/f ${\cal U}$ is associative. \hfill $\blacksquare$ \renewcommand{\theequation}{{\rm A}3.\arabic{equation}} \setcounter{equation}{0} \subsection*{Appendix 3. Lie Algebras of Vector Fields on Lie Groups} Formula \begin{equation} X * Y = XY', \qquad X, Y \in C^\infty (S^{-1}), \quad {}' = {d \over dz}, \end{equation} provides a quasiassociative structure on the Lie algebra of vector f\/ields on the circle, ${\cal D}(S^1)$. Formula~[5] \begin{equation} (X * Y)^i = \sum_s X^s Y^i,_{s} \end{equation} provides a quasiassociative structure on the Lie algebra of vector f\/ields on ${\mathbf R}^n, {\cal D} ({\mathbf R}^n)$. This suggests that for some manifolds, similar structure exists for their Lie algebras of vector f\/ields. (This will be proven below for $GL(n, {\mathbf R})$ and $GL(n, {\mathbf C})$.) The parallelizable manifolds are the simplest, and Lie groups are simpler still. \medskip \noindent {\bf Proposition A3.3.} {\it Let ${\cal R}$ be a finite-dimensional quasiassociative algebra over ${\mathbf R}$, ${\cal G} = Lie ({\cal R})$, and $G$ a connected Lie group with the Lie algebra ${\cal G}$. Then the Lie algebra of vector fields on $G$, ${\cal D} (G)$, has a quasiassociative representation.} \medskip \setcounter{equation}{3} \noindent {\bf Proof.} Let $(e_i)$ be a basis in ${\cal R}$. Then \begin{equation} e_i e_j = \sum_s \theta^s_{ij} e_s, \end{equation} with some structure constants $\theta^s_{ij} \in {\mathbf R}$. The quasiassociativity condition \begin{equation} (e_i e_j) e_\kappa - e_i (e_j e_\kappa) = (e_j e_i) e_\kappa - e_j (e_i e_\kappa), \qquad \forall \; i, j, \kappa, \end{equation} translates into the equality \begin{equation} \sum_s \left(\theta^s_{jk} \theta^r_{is} - \theta^s_{ij} \theta^r_{sk}\right) = \sum_s \left(\theta^s_{ik} \theta^r_{js} - \theta^s_{ji} \theta^r_{sk}\right), \end{equation} or \begin{equation} \sum_s \left(\theta^s_{jk} \theta^r_{is} - \theta^s_{ik} \theta^r_{js}\right)= \sum_s c^s_{ij} \theta^r_{sk}, \end{equation} where \begin{equation} c^s_{ij} = \theta^s_{ij} - \theta^s_{ji} \end{equation} are the structure constants of the Lie algebra ${\cal G} = Lie ({\cal R})$: \begin{equation} [e_i, e_j] = e_i e_j - e_j e_i = \sum_s c^s_{ij} e_s. \end{equation} Denote by $\hat e_i$ the left-invariant vector f\/ields on $G$ generated by the elements $e_i \in {\cal G},$ so that \begin{equation} \hat e_i \hat e_j - \hat e_j \hat e_i = \sum_j c^s_{ij} \hat e_s. \end{equation} In this basis, every vector f\/ield on $G$ can be identif\/ied with a vector from $C^\infty (G)^{\mbox{\scriptsize \rm dim}(G)}$: \begin{equation} X \in {\cal D} (G) \ \Rightarrow \ X = \sum_i X^i \hat e_i, \qquad X^i \in C^\infty (G). \end{equation} For $X = \sum X^i \hat e_i$, $Y = \sum Y^j \hat e_j \in {\cal D} (G)$, set \begin{equation} (X * Y)^r = \sum_\alpha X^\alpha \hat e_\alpha (Y^r) + \sum_{\alpha \beta} X^\alpha Y^\beta \theta^r_{\alpha \beta}. \end{equation} We are going to show that this multiplication makes ${\cal D} (G) \approx C^\infty (G)^{\mbox{\scriptsize \rm dim}(G)}$ into a quasiassociative algebra; the boundary conditions are satisf\/ied since \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle \sum_r (X * Y - Y * X)^r \hat e_r = \sum_{\alpha r} \left[X^\alpha \hat e_\alpha (Y^r) - Y^\alpha \hat e_\alpha (X^r) \right] \hat e_r \vspace{3mm}\\ \displaystyle \qquad + \sum_{\alpha \beta r} X^\alpha Y^\beta c^r_{\alpha \beta} \hat e_r = \sum_{\alpha \beta} \left[X^\alpha \hat e_\alpha, Y^\beta \hat e_\beta\right] =[X, Y]. \end{array} \end{equation} Now, \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle (X (YZ))^r = \sum_\alpha X^\alpha \hat e_\alpha ((YZ)^r) + \sum_{\alpha \beta} \theta^r_{\alpha \beta} X^\alpha (YZ)^\beta \vspace{3mm}\\ \displaystyle \qquad = \sum_\alpha X^\alpha \hat e_\alpha \left(\sum_\mu Y^\mu \hat e_\mu (Z^r)+ \sum_{\mu \nu} \theta^r_{\mu \nu} Y^\mu Z^\nu\right) \vspace{3mm}\\ \displaystyle \qquad + \sum_{\alpha \beta} \theta^r_{\alpha \beta} X^\alpha \left(\sum_\mu Y^\mu \hat e_\mu (Z^\beta) + \sum_{\mu \nu} \theta^\beta_{\mu \nu} Y^\mu Z^\nu\right) \vspace{3mm}\\ \displaystyle \qquad = \sum_{\alpha \mu} X^\alpha \hat e_\alpha (Y^\mu) \hat e_\mu (Z^r) + \sum_{\alpha \mu} X^\alpha Y^\mu \hat e_\alpha \hat e_\mu (Z^r) \vspace{3mm}\\ \displaystyle \qquad + \sum_{\alpha \mu \nu} X^\alpha \theta^r_{\mu \nu} \left(\hat e_\alpha (Y^\mu) Z^\nu + Y^\mu \hat e_\alpha (Z^\nu)\right) \vspace{3mm}\\ \displaystyle \qquad +\sum_{\alpha \mu \nu} \theta_{\alpha \nu}^r X^\alpha Y^\mu \hat e_\mu (Z^\nu) + \sum_{\alpha s \mu \nu} \theta^r_{\alpha s} \theta^s_{\mu \nu} X^\alpha Y^\mu Z^\nu, \end{array} \end{equation} \begin{equation} \hspace*{-5pt}\begin{array}{l} \displaystyle ((X Y)Z)^r = \sum_\mu (XY)^\mu \hat e_\mu (Z^r) + \sum_{s \nu} \theta^r_{s \nu} (XY)^s Z^\nu = \sum_{\mu \alpha} X^\alpha \hat e_\alpha (Y^\mu) \hat e_\mu (Z^r) \vspace{3mm}\\ \displaystyle \qquad +\sum_{\mu \alpha \beta} \theta^\mu_{\alpha \beta} X^\alpha Y^\beta \hat e_\mu (Z^r) + \sum_{s \nu} \theta^r_{s \nu} \left(\sum_\alpha X^\alpha \hat e_\alpha (Y^s)+ \sum_{\alpha \beta} \theta^s_{\alpha \beta} X^\alpha Y^\beta \right) Z^\nu. \end{array} \end{equation} Thus, \[ (X(YZ) - (XY) Z)^r \] \renewcommand{\theequation}{{\rm A}3.\arabic{equation}{\rm a}} \setcounter{equation}{15} \begin{equation} \qquad = \sum_{\alpha \mu} X^\alpha Y^\mu \hat e_\alpha \hat e_\mu (Z^r) \end{equation} \renewcommand{\theequation}{{\rm A}3.\arabic{equation}{\rm b}} \setcounter{equation}{15} \begin{equation} \qquad + \sum_{\alpha \mu \nu} X^\alpha Y^\mu \theta^r_{\mu \nu} \hat e_\alpha (Z^\nu) + \sum_{\alpha \mu \nu} X^\alpha Y^\mu \theta^r_{\alpha \nu} \hat e_\mu (Z^\nu) \end{equation} \renewcommand{\theequation}{{\rm A}3.\arabic{equation}{\rm c}} \setcounter{equation}{15} \begin{equation} \qquad + \sum_{\alpha s \mu \nu} X^\alpha Y^\mu Z^\nu \left(\theta^r_{\alpha s} \theta^s_{\mu \nu} - \theta^r_{s \nu} \theta^s_{\alpha \mu}\right) \end{equation} \renewcommand{\theequation}{{\rm A}3.\arabic{equation}{\rm d}} \setcounter{equation}{15} \begin{equation} \qquad - \sum_{\mu \alpha \beta} \theta^\mu_{\alpha \beta} X^\alpha Y^\beta \hat e_\mu (Z^r). \end{equation} Interchanging $X$ and $Y$, subtracting the resulting expressions, noticing that (A3.16b) is symmpletric in $(X, Y)$, and using formulae (A3.10,8), we arrive at the following identity to be verif\/ied: $0 {\mathop{=}\limits^{?}}$ \renewcommand{\theequation}{{\rm A}3.\arabic{equation}{\rm a}} \setcounter{equation}{16} \begin{equation} =\sum_{\alpha \mu s} X^\alpha Y^\mu \left(\theta^s_{\alpha \mu} - \theta^s_{\mu\alpha}\right) \hat e_s (Z^r) \end{equation} \renewcommand{\theequation}{{\rm A}3.\arabic{equation}{\rm b}} \setcounter{equation}{16} \begin{equation} + \sum_{\alpha s \mu \nu} X^\alpha Y^\mu Z^\nu \left(\theta^r_{\alpha s} \theta^s_{\mu\nu} - \theta^v_{s \nu} \theta^s_{\alpha \mu} - \theta^r_{\mu s} \theta^s_{\alpha \nu} + \theta^r_{s \nu} \theta^s_{\mu \alpha} \right) \end{equation} \renewcommand{\theequation}{{\rm A}3.\arabic{equation}{\rm c}} \setcounter{equation}{16} \begin{equation} - \sum_{\mu \alpha \beta} \left(\theta^\mu_{\alpha \beta} - \theta^\mu_{\beta\alpha} \right) X^\alpha Y^\beta \hat e_\mu (Z^r). \end{equation} The expressions (A3.17a) and (A3.17c) cancel each other out. The sum (A3.17b) vanishes due to the quasiassociativity condition (A3.6). \hfill $\blacksquare$
\section{Introduction} Cold dark matter (CDM) cosmologies predict that the first baryonic objects appeared near the Jeans mass ($\sim 10^6~{\rm M_\odot}$) at redshifts as high as $z\sim30$, and larger objects assembled later (Haiman \& Loeb 1999b, and references therein). It is natural to identify these objects as the sites where the first stars or quasar black holes formed. Observationally, bright quasars are currently detected out to $z\sim 5$ (Fan et al. 1999), and galaxies out to $z\sim 6.7$ (Chen et al. 1999; Weymann et al. 1998). Although the abundance of optically and radio bright quasars declines at $z\gsim 2.5$ (Schmidt et al. 1995; Pei 1995; Shaver et al. 1996), simple models based on the Press--Schechter formalism reproduce this decline and simultaneously predict a population of low-luminosity quasars that are too faint to be detected in current surveys (Haiman \& Loeb 1998; Haehnelt et al. 1998). Preliminary evidence for such a population might already be indicated by the X--ray luminosity function (LF) of quasars recently measured by ROSAT (Miyaji et al. 1998a). The X--ray data probes fainter quasars than the optical data does, and has not revealed a decline in the abundance of high-redshift quasars as found in the optical (e.g. Schmidt et al. 1995). The current census of high-redshift galaxies provides an estimate of the global evolution of the star formation rate (SFR) in the redshift range $0\leq z\lsim 5$ (Madau 1999). Nevertheless, there is still considerable uncertainty about the evolution of the SFR at the highest redshifts probed ($z\gsim 3$). Recent sub--mm observations from SCUBA indicate the existence of a substantial population of dust--obscured star--forming galaxies that could raise the implied SFR at high redshift (Barger et al. 1999, and references therein). This uncertainty is particularly important, since it leaves open the question of whether the observed population of stars produce sufficient ionizing radiation by $z\sim5$ to reionize the intergalactic medium (IGM). The indication from the present data is that the SFR either declines or remains roughly constant at $z\gsim 3$; however, in order to satisfy the Gunn--Peterson constraint for the ionization of the IGM, the SFR needs to rise at high redshifts above its value at $z\sim 3$ (Madau 1999). The presence of a substantial population of faint quasars at $z\gsim 5$ is therefore consistent with current data, and is suggested by cosmological models for hierarchical structure formation. If future data would indicate a non-increasing cosmic SFR at $z\ga 3$, then quasars would be the likely sources of reionization. Recent theoretical models have focused on the properties of the quasar population in the optical (Haehnelt \& Rees 1993; Haiman \& Loeb 1998; Haehnelt et al. 1998), and in the infrared (e.g. Sanders 1999; Almaini et al.~1999). These models were shown to be only mildly constrained by the lack of $z\gsim 3.5$ quasars in the Hubble Deep Field (HDF; see Haiman, Madau \& Loeb 1999). In this {\it Letter} we point out that quasars can be best distinguished from star forming galaxies at high redshifts by their X-ray emission. First, we illustrate the agreement between our simplest hierarchical model for quasars and existing ROSAT data on the X--ray LF (Miyaji et al. 1998a) and the soft X--ray background (XRB, Miyaji et al. 1998b). We then show that forthcoming X-ray observations with the {\it Chandra X-ray Observatory} ({\it CXO}; formerly known as {\it AXAF}) will be able to probe the abundance of quasars during the reionization of the Universe at $z\ga 5$. \section{Model Description and Quasar Spectra} Our model is based on the Press--Schechter mass function of CDM halos (see Haiman \& Loeb 1998 for complete details). We assume that each halo of mass $M_{\rm halo}$ forms a central black hole of mass $M_{\rm bh}$. The black hole mass fraction $r\equiv M_{\rm bh}/M_{\rm halo}$ is assumed to obey a log-Gaussian probability distribution \beq p(r)=\exp [-(\log r - \log r_0)^2/2\sigma^2] \label{eq:scat} \eeq with $\log r_0=-3.5$ and $\sigma=0.5$ (Haiman \& Loeb 1999b) These values roughly reflect the distribution of black hole to bulge mass ratios found in a sample of 36 local galaxies (Magorrian et al. 1998) for a baryonic mass fraction of $\sim (\Omega_{\rm b}/\Omega_0)\approx 0.1$. We further postulate that each black hole emits a time--dependent bolometric luminosity in proportion to its mass, $L_{\rm q}\equiv M_{\rm bh}f_{\rm q}=M_{\rm bh}L_{\rm Edd} \exp(-t/t_0)$, where $L_{\rm Edd}=1.5\times10^{38}~M_{\rm bh}/{\rm M_\odot}~{\rm erg~s^{-1}}$ is the Eddington luminosity, $t$ is the time elapsed since the formation of the black hole, and $t_0=10^6$yr is the characteristic quasar lifetime. Finally, we assume that the shape of the emitted spectrum follows the mean spectrum of the quasar sample in Elvis et al. (1994) up to a photon energy of 10 keV. We extrapolate the spectrum up to $\sim 50$ keV, assuming a spectral slope of $\alpha$=0 (or a photon index of -1). \vspace{+0.2cm} \begin{figurehere} \plotone{fig1.ps}\vspace{-0.4cm} \caption[Spectra] {\label{fig:spectrum} \footnotesize The observed spectra of quasars with a central black hole mass of $M_{\rm bh}=10^8~{\rm M_\odot}$. The upper curves correspond to a source redshift of $z_{\rm s}=6$ and the lower curves to a source redshift of $z_{\rm s}=11$. In both cases we assume sudden reionization at $z_{\rm r}=10$. The dashed curves show the intrinsic spectral shape, taken from Elvis et al. (1994). We include the opacity of the neutral H and He in the IGM at $z>z_{\rm r}$ and the opacity due to electron scattering and the (extrapolated) Ly$\alpha$ forest in the IGM at $z<z_{\rm r}$.} \end{figurehere} \vspace{0.4cm} This simple model was demonstrated to accurately reproduce the evolution of the optical luminosity function in the B--band (Pei 1995) at redshifts $z\gsim 2.2$ (Haiman \& Loeb 1998). Because our model incorporates several simplifying assumptions, we regard it as the minimal toy model which successfully reproduces the existing data. If one of our input assumptions was drastically violated and our model had failed to fit the observed LF, then a modification of its basic ingredients would be needed. In this {\it Letter}, we focus on the predictions of this minimal model in anticipation of the forthcoming launch of {\it CXO}; an investigation of a broader range of plausible toy models will be made elsewhere. We adopt the concordance cosmology of Ostriker \& Steinhardt (1995), namely a $\Lambda$CDM model with a tilted power spectrum ($\Omega_0,\Omega_\Lambda, \Omega_{\rm b},h,\sigma_{8h^{-1}},n$)=(0.35, 0.65, 0.04, 0.65, 0.87, 0.96). Figure~\ref{fig:spectrum} shows the adopted spectrum of quasars, assuming a black hole mass $M_{\rm bh}=10^8{\rm M_\odot}$, placed at two different redshifts, $z_{\rm s}=11$ and $z_{\rm s}=6$. In computing the intergalactic absorption, we included the opacity of both hydrogen and helium as well as the effect of electron scattering. We assumed that reionization occurred at $z_{\rm r}=10$ and that at higher redshifts the IGM was homogeneous and fully neutral. At lower redshifts, $0<z<z_{\rm r}$, we included the hydrogen opacity of the Ly$\alpha$ forest given by Madau (1995), and extrapolated his fitting formulae for the evolution of the number density of absorbers beyond $z=5$ when necessary. As Figure~\ref{fig:spectrum} shows, the minimum black hole mass detectable by the $\sim 2\times 10^{-16}~{\rm erg~s^{-1}~cm^{-2}}$ flux limit of {\it CXO} (see below) is $M_{\rm bh}\sim 10^8~{\rm M_\odot}$ at $z=10$ and $M_{\rm bh}\sim 2\times10^7~{\rm M_\odot}$ at $z=5$. In our model, the corresponding halo masses are $M_{\rm halo}\sim 3\times10^{11}~{\rm M_\odot}$, and $M_{\rm halo}\sim 6\times 10^{10}~{\rm M_\odot}$, respectively. Although such massive halos are rare, their abundance is detectable in wide-field surveys. Note that an accurate determination of the spectrum below $\sim0.1$keV could have provided an estimate of the reionization redshift $z_{\rm r}$. Unfortunately, this spectral regime suffers from Galactic absorption (O'Flaherty and Jakobsen 1997) and is outside the 0.4--6keV detection band of {\it CXO}. \section{The X--ray Luminosity Function} In our model, the X--ray luminosity function at a redshift $z$ [in ${\rm Mpc^{-3}~(erg/s)^{-1}}$] is given by a sum over halos that formed just before that redshift, \beq \phi(L_X,z) =\int_{0}^{t(z)} \frac{dt}{f_{\rm X}f_{\rm q}(\Delta t)} \left.\frac{d^2N}{dM_{\rm bh}dt}\right|_{M_{\rm bh}= \frac{L_X}{f_{\rm X}f_{\rm q}(\Delta t)}} \,\,\, \label{eq:LF} \eeq where $L_X$ is the observed X--ray luminosity in the instrument's detection band ($0.5$--$3$ keV for ROSAT and 0.4--6 keV for {\it CXO}); $f_{\rm X}$ is the fraction of the quasar's bolometric luminosity emitted in this band; $d^2N/dM_{\rm bh}dt$ is the black halo formation rate, given by a convolution of the Press--Schechter halo mass function with equation~(\ref{eq:scat}); and $\Delta t=t(z)-t$ is the time elapsed from a cosmic time $t$ until a redshift $z$. Although our model was constructed so as to fit the observed optical LF, Figure~\ref{fig:LF} demonstrates that it is also in good agreement with the data on the X--ray LF. This implies that the choice of quasar spectrum in our model is reasonable. The solid curve in this figure shows the prediction of equation~(\ref{eq:LF}) at $z=3.5$, near the highest redshift where X-ray data is available. The bottom curve corresponds to a cutoff in circular velocity for the host halos of $v_{\rm circ}\geq 50~{\rm km~s^{-1}}$, which is introduced here in order not to over--predict the number of quasars in the HDF (Haiman, Madau \& Loeb 1999). The data points are from recent ROSAT measurements, and the dashed curve in this figure shows a fitting formula from Miyaji et al. (1998). Note that the faintest quasar actually observed has $L_X\lsim 10^{44}~{\rm erg~s^{-1}}$, and the fitting formula below this luminosity is highly uncertain. Nevertheless, it is important to remember that our model was calibrated so as to fit the observed optical luminosity function of quasars; the existence of a population of obscured quasars which are faint in the optical band but bright in X--rays could increase the number counts beyond our model predictions. \vspace{+0.2cm} \begin{figurehere} \plotone{fig2.ps}\vspace{-0.4cm} \caption[Luminosity Function at z=3.5] {\label{fig:LF} \footnotesize The predicted X--ray luminosity function at $z=3.5$ in our model (solid curves). The lower curve shows the effect of a cutoff in circular velocity for the host halos of $v_{\rm circ}\geq 50~{\rm km~s^{-1}}$. The ROSAT data points are adopted from Miyaji et al. (1998a) and the dashed curve shows their fitting formula (for our background cosmology). } \end{figurehere} \vspace{0.1cm} \section{The X--ray Background} Existing estimates of the X--ray background (XRB) provide another useful check on our quasar model. The unresolved background flux at a photon energy $E$ is given by (Peebles 1993) \beq F(E)= c \int_0^\infty dz \frac{dt}{dz} j(E_z,z)\,\,\,\,\,{\rm keV~cm^{-2}~s^{-1}~sr^{-1}~keV^{-1}}, \label{eq:xrb} \eeq where $E_z=E(1+z)$; and $j(E_z,z)$ is the comoving emissivity at a local photon energy $E_z$, in units of ${\rm keV~cm^{-3}~s^{-1}~sr^{-1}~keV^{-1}}$, from quasars shining at a redshift $z$. This emissivity is a sum over all quasars whose individual observed flux at $z=0$ is below the ROSAT PSPC detection limit for discrete sources of $2\times10^{-15}~{\rm erg~cm^{-2}~s^{-1}}$ (Hasinger \& Zamorani 1997). Figure~\ref{fig:xrb} shows the predicted spectrum of the XRB in our model at $z=0$ (solid lines). In computing the background spectrum, we ignored the HI absorption in the IGM, since it is negligible at energies above 100 eV. We also carried out the integral in equation~(\ref{eq:xrb}) only for $z>2$, the redshift range where our model is valid (Haiman \& Loeb 1998). The short dashed lines show the predicted fluxes assuming a steeper spectral slope beyond 10 keV ($\alpha=-0.5$, or a photon index of -1.5). The long dashed line shows the 25\% unresolved fraction of the soft XRB observed with ROSAT (Miyaji et al. 1998b; Fabian \& Barcons 1992). This fraction represents the observational upper limit on the component of the soft XRB that could in principle arise from high-redshift quasars. As the figure shows, our quasar model predicts an unresolved flux just below this limit in the 0.5-3 keV range. The model also predicts that most ($\gsim 90\%$) of this yet unresolved fraction arises from quasars beyond $z=5$. \vspace{+0.2cm} \begin{figurehere} \plotone{fig3.ps}\vspace{-0.4cm} \caption[X--Ray Background] {\label{fig:xrb} \footnotesize Spectrum of the soft X--ray background in our model. We assume that the median X-ray spectrum of each source follows the mean spectrum of Elvis et al. (1994) up to 10 keV, and has a spectral slope of 0.5 (solid lines) or -0.5 (short--dashed lines) at higher photon energies. The lower curves show the spectra resulting from quasars with redshifts between $2<z<5$, and the upper curves include contributions from all redshifts $z>2$. In both cases, we include only quasars whose individual fluxes are below the ROSAT PSPC limit of $2\times10^{-15}~{\rm erg~cm^{-2}~s^{-1}}$ for the detection of a discrete source. The long--dashed line shows the unresolved fraction (assumed to be 25\%) of the soft X--ray background spectrum from Miyaji et al. (1998b).} \end{figurehere} \vspace{0.1cm} \section{Predicted Number Counts for the Chandra X-ray Observatory} By summing the luminosity function over redshifts, we obtain the number counts of quasars per solid angle expected to be detectable in a flux interval $dF_X$ around $F_X=L_X/4\pi d_L^2(z)$, for all sources above a redshift $z$: \beq \frac{dN}{dF_X d\Omega} = \int_z^\infty dz \left(\frac{dV}{dz d\Omega}\right) \phi(L_X,z) 4\pi d_L^2(z), \label{eq:counts} \eeq where $(dV/dzd\Omega)$ is the comoving volume element per unit redshift and solid angle, and $d_L(z)$ is the luminosity distance at a redshift $z$. In Figure~\ref{fig:counts}, we show the predicted counts from equation~(\ref{eq:counts}) in the 0.4--6keV energy band of the CCD Imaging Spectrometer (ACIS) of {\it CXO}. Note that these curves are insensitive to our extrapolation of the template spectrum beyond 10 keV. The figure is normalized to the $17^\prime \times 17^\prime$ field of view of the imaging chips. The solid curves show that of order a hundred quasars with $z>5$ are expected per field. The abundance of quasars at higher redshifts declines rapidly; however, a few objects per field are still detectable at $z\sim 8$. The dashed lines show the results for a minimum circular velocity of the host halos of $v_{\rm circ}\geq 100~{\rm km~s^{-1}}$, and imply that the model predictions for the {\it CXO} satellite are not sensitive to such a change in the host velocity cutoff. This is because the halos shining at the {\it CXO} detection threshold are relatively massive, $M_{\rm halo}\sim 10^{11}~{\rm M_\odot}$, and possess a circular velocity above the cutoff. In principle, the number of predicted sources would be lower if we had assumed a steeper spectral slope. However, the agreement between the LF predicted by our model at $z\approx 3.5$ and that inferred from ROSAT observations would be upset by such a change, and require a modification of the model that would in turn tend to counter-balance the decrease in the predicted counts. \vspace{+0.2cm} \begin{figurehere} \plotone{fig4.ps}\vspace{-0.4cm} \caption {\label{fig:counts} \footnotesize The total number of quasars with redshift exceeding $z=5$, $z=7$, and $z=10$ are shown as a function of observed X-ray flux in the {\it CXO} detection band. The solid curves correspond to a cutoff in circular velocity for the host halos of $v_{\rm circ}\geq 50~{\rm km~s^{-1}}$, the dashed curves to a cutoff of $v_{\rm circ}\geq 100~{\rm km~s^{-1}}$. The vertical dashed line show the {\it CXO} sensitivity for a 5$\sigma$ detection of a point source in an integration time of $5\times10^5$ seconds.} \end{figurehere} \vspace{0.1cm} \section{Discussion} The existence of quasars at redshifts $z\ga 5$ has important consequences for the reionization history of the Universe. Quasars can be easily distinguished from stellar systems by their X-ray emission. We have demonstrated that state--of--the--art X-ray observations could provide more stringent constraints on quasar models than currently provided by the Hubble Deep Field (Haiman, Madau, \& Loeb 1999). In particular, we have found that forthcoming X--ray observations with the {\it CXO} satellite might detect of order a hundred quasars per field of view in the redshift interval $5\la z\la 8$. Our numerical estimates are based on the simplest toy model for quasar formation in a hierarchical CDM cosmology, that satisfies all the current observational constraints on the optical and X-ray luminosity functions of quasars. Although a more detailed analysis is needed in order to assess the modeling uncertainties in our predictions, the importance of related observational programs with {\it CXO} is evident already from the present analysis. Other future instruments, such as the HRC or the ACIS-S cameras on {\it CXO}, or the EPIC camera on {\it XMM}, might also be useful in searching for high--redshift quasars. Follow-up optical and infrared observations are needed in order to identify the redshifts of the faint point-like sources that might be detected by the {\it CXO} satellite. Quasars emit a broad spectrum which extends into the UV and includes strong emission lines, such as Ly$\alpha$. For quasars near the {\it CXO} detection threshold, the fluxes at $\sim 1\mu$m are expected to be relatively high, $\sim 0.5$--$0.8~\mu$Jy. Therefore, infrared spectroscopy of X--ray selected quasars could prove to be a particularly useful approach for unraveling the reionization history of the intergalactic medium at $z\ga 5$ (for the potential lessons to be learned, see Miralda-Escud\'e 1998; Haiman \& Loeb 1999a; Loeb \& Rybicki 1999; and Rybicki \& Loeb 1999). At present, the best constraints on hierarchical models of the formation and evolution of quasars originate from the Hubble Deep Field. However, {\it HST} observations are only sensitive to a limiting magnitude of $V\sim29$ and cannot probe the earliest quasars. The Next Generation Space Telescope ({\it NGST}), scheduled for launch in 2008, will achieve nJy sensitivity at wavelengths $1-3~\mu{\rm m}$, and could directly probe the earliest quasars (Haiman \& Loeb 1998). The combination of X-ray data from the {\it CXO} satellite and infrared spectroscopy from {\it NGST} could potentially resolve one of the most important open questions about the thermal history of the Universe, namely whether the intergalactic medium was reionized by stars or by accreting black holes. \acknowledgments We thank G. Hasinger and R. Mushotzky for discussions that motivated this study. We also thank N. White and M. Elvis for useful comments, and T. Miyaji for supplying the data in Figure~\ref{fig:LF}. ZH was supported by the DOE and the NASA grant NAG 5-7092 at Fermilab. AL was supported in part by NASA grants NAG 5-7039 and NAG 5-7768.
\section{Introduction} It is a well known fact that anomaly cancellation in $SO(32)$ Type I string may be understood as a direct consequence of the cancellation of tadpoles of Ramond-Ramond fields. In fact the implication runs in both directions and anomaly cancellation implies also tadpole cancellation. This is not so surprising since in $D=10$ the anomaly cancellation constraints fix almost uniquely the massless particle content of the theory. In Type I vacua in lower dimensions, like in Type IIB $D=4,6$ orientifolds, tadpole cancellation constraints do also imply anomaly cancellation. An interesting question arises regarding the extent up to which anomaly cancellation and tadpole cancellation are still equivalent in lower dimensions. One of the purposes of this paper is to address this question. We concentrate in the study of $D=4$, $N=1$ Type IIB compact orientifolds, although analogous results are obtained in $D=6$. We find that tadpole cancellation is in general a more stringent constraint than anomaly cancellation in the case of $D=4$ orientifolds. In order to show this, we rewrite the anomaly cancellation conditions in terms of traces of Chan-Paton twist matrices acting on 9-branes and/or 5-branes. In this way, it is shown that in general only a subset of the complete tadpole cancellation conditions is recovered. Thus, there are certain tadpole cancellation constraints which are actually {\it not} required for anomaly cancellation. Which those are depends on the structure of the twist group of the orientifold and also on the simultaneous presence or not of 9-branes and 5-branes\footnote{We are referring here to chiral $D=4$ orientifolds. Non-chiral ones like the $Z_2\times Z_2$ model of ref.\cite{bl} obviously do not get any constraint from anomaly cancellation.} This procedure allows a broad scan for solutions of tadpole cancellation conditions in $D=4,6$ $N=1$, Type IIB orientifolds. In particular, we find that there are certain $D=4$ and $D=6$ orientifolds which admit many more solutions than previously reported in the literature. The relationship between tadpole cancellation and anomaly cancellation was considered in ref.\cite {lr} \footnote{See also \cite{bi1,bi2,pu} for analogous results in six-dimensional orientifolds.} in the context of $D=4$, $N=1$ gauge theories on the world-volume of D3-branes sitting at $Z_N\times Z_M$ singularities \footnote{$N=1$ supersymmetric theories form D3 branes at orbifold singularities have been studied in \cite{dgm, ks,lnv,hu,hh}. The inclusion of orientifold projections has been discussed in \cite{lykken,kak1,kak2,iru}.}. In that reference an equivalence between tadpole and anomaly cancellation conditions was found. Those models differ from the class we consider in several important respects. In particular they are non-compact models (the six dimensions transverse to the D3-branes are not compactified) and in addition they only contain D3-branes. In this paper we consider {\it compact} orientifolds and this fact makes the following important difference. Tadpoles are sources for the RR potentials of the theory. In non-compact models some of the twisted RR fields can propagate on non-compact directions and carry the RR-flux off to infinity. The models are consistent without imposing cancellation of the corresponding twisted tadpoles. In the compact models we are to consider, the RR-flux cannot escape to infinity but is trapped in the compact space. Thus additional constraints may appear. A second difference is that we consider the simultaneous presence of two types of D-branes, 9-branes and 5-branes. This is equivalent by T-duality to considering both 3-branes and 7-branes in the context of \cite{lr}. However, in this reference the emphasis is in the gauge theory on the world-volume of the D3-branes, while the D7-branes are considered non-dynamical. This is justified since the D7-branes have more non-compact dimensions than the D3-branes. In our compact models, however, D7 branes wrap compact direction and yield truly four-dimensional fields. Thus the cancellation of anomalies from gauge groups living on the D7-branes lead to additional constraints, not present in the non-compact models. The considerations in this and the preceding paragraph explain the difference between our conclusions and those for non-compact models in \cite{lr}. The outline of the paper is as follows. In the next section we review some facts about compact $D=4$, $N=1$ Type IIB orientifolds and establish the notation needed for the remaining sections. In section 3.1 we study the compact orientifolds without even order generators i.e., $Z_3$, $Z_7$ and $Z_3\times Z_3$ orientifolds. In this case the models have only D9-branes and tadpole conditions uniquely fix the gauge group. However in the $Z_3\times Z_3$ case it is shown how tadpole cancellation conditions are stronger than anomaly cancellation conditions. In section 3.2 we study the compact orientifolds with even order twists. In this case both D9-branes and D5-branes are present. The $Z_4$, $Z_8$, $Z_8'$ and $Z_{12}'$ orientifolds with standard orientifold projection are shown to be necessarily anomalous, in agreement with the results of ref. \cite{afiv} in which they were shown to have non-vanishing tadpoles. We present other examples (the $Z_{12}$ orientifold) in which tadpole conditions are explicitly shown to be stronger than anomaly cancellation ones. New solutions for the tadpole conditions are shown for the $Z_6$ and $Z_{12}$ orientifolds leading to a variety of gauge groups previously overlooked in the literature. We study the cancellation of $U(1)$ anomalies in section~4. We find that cancellation of non-Abelian anomalies guarantees the cancellation of Abelian anomalies. In addition it is shown that in the $D=4$ compact orientifold case cancellation of non-Abelian anomalies also implies the cancellation of mixed $U(1)$-gravitational anomalies. Section 5 is left for some final comments and conclusions. In particular we compare our results to those found for non-compact orientifolds and discuss the origin of the non-equivalence of tadpole/anomaly conditions. We also briefly discuss the equivalent results for $D=6$, $N=1$ compact IIB orientifolds. \section{$D=4$, $N=1$, Type IIB Orientifolds} \label{basics} In this section we summarize the basic ingredients \cite {sagnotti,bs,gp,afiv} and notation needed in the construction of $D=4,N=1$ orientifold. The reader is referred to \cite{afiv} for further details. In a Type IIB orientifold, the toroidally compactified theory is divided out by the joint action of a discrete symmetry group $G_1$, (like $Z_N$ or $Z_N\times Z_M$) together with a world sheet parity operation $\Omega$, exchanging left and right movers. $\Omega $ action can be accompanied by extra operations thus leading to generic orientifold group $G_1+ {\Omega} G_2$ with ${\Omega}h {\Omega} h' \in G_1$ for $h,h' \in G_2$ In this article we will refer to the cases $G_1=G_2$ and $G_1=Z_N$ or $G_1=Z_N\times Z_M$ and such that $D=4$ $N=1$ theories are obtained, when the twist $\Omega$ is performed on Type IIB compactified on $T^6/G_1$. The allowed orbifold groups, acting crystalographically on $T^6$ leading to $N=1$ unbroken supersymmetry were classified in \cite{dhvw}. The list, with corresponding twist vector eigenvalues $v=(v_1,v_2,v_3)$ associated to the $Z_N$ orbifold twist $\theta$ is given in Table~\ref{tzn}. \begin{table}[htb] \renewcommand{\arraystretch}{1.25} \begin{center} \begin{tabular}{|c|c||c|c||c|c|} \hline $Z_3$ & $\frac13(1,1,-2)$ & $Z_6^{\prime}$ & $\frac16(1,-3,2)$ & $Z_8^{\prime}$ & $\frac18(1,-3,2)$ \\ $Z_4$ & $\frac14(1,1,-2)$ & $Z_7$ & $\frac17(1,2,-3)$ & $Z_{12}$ & $\frac1{12}(1,-5,4)$ \\ $Z_6$ & $\frac16(1,1,-2)$ & $Z_8$ & $\frac18(1,3,-4)$ & $Z_{12}^{\prime}$ & $\frac1{12}(1,5,-6)$ \\ \hline \end{tabular} \end{center} \caption{$Z_N$ actions in \Deq4.} \label{tzn} \end{table} Orientifolding closed Type IIB string introduces a Klein-bottle unoriented world-sheet. Amplitudes on such a surface contain tadpole divergences. In order to eliminate such unphysical divergences Dp-branes must be generically introduced. In this way, divergences occurring in the open string sector cancel up the closed sector ones and produce a consistent theory. Tadpole cancellation is interpreted as cancellation of the charge carried by RR form potentials. For $Z_N$, with $N$ odd, only D9-branes are required. They fill the full space-time and six dimensional compact space. For $N$ even, D$5_k$-branes, with world-volume filling space-time and the $k^{th}$ complex plane, may be required. This is so whenever the orientifold group contains the element $\Omega R_i R_j$, for $k\neq i,j$. Here $R_i$ ($R_j$) is an order two twist of the $i^{th}$ ($j^{th}$) complex plane. In what follows we consider cases with only one set of five branes. $Z_N$ twists in Table \ref{tzn} were organized in such a way that, for even $N$, the order two element $R=\theta^{N/2}$ inverts the complex planes $Y_1$ and $Y_2$ and thus the corresponding orientifolds have D$5_3$-branes, filling space-time and compact dimension $Y_3$. Open string states are denoted by $|\Psi, ab \rangle $, where $\Psi$ refers to world-sheet degrees of freedom while the $a,b$ Chan-Paton indices are associated to the open string endpoints lying on D$p$-branes and D$q$-branes respectively. These Chan-Paton labels must be contracted with a hermitian matrix $\lambda ^{pq} _{ab}$. The action of an element of the orientifold group on Chan-Paton factors is achieved by a unitary matrix $\gamma _{g,p}$ such that $g: \lambda ^{pg} \rightarrow \gamma _{g,p} \lambda ^{pq} \gamma^{-1}_{g,q}$. We denote by $\gamma _{k,p}$ the matrix associated to the $Z_N$ orbifold twist $\theta ^k $ acting on a Dp-brane. Consistency under group transformations imposes restrictions on the representations $\gamma _g$. For instance, from $\Omega ^2 =1$ it follows that \begin{equation} \gamma_{\Omega,p }=\pm\gamma^T_{\Omega,p } \label{gomt} \end{equation} Tadpole cancellation imposes further constraints on $\gamma _g$. Since we are planning to compare such restrictions with those coming from anomaly cancellation of gauge theories on D5 and D9-brane configurations, we will not impose the former in what follows, but just consider generic actions obeying the algebraic consistency conditions. Nevertheless, we will perform a definite choice of signs in (\ref{gomt}), namely \begin{eqnarray} \gamma_{\Omega,9 } & = & \gamma^T_{\Omega,9 } \nonumber \\ \gamma_{\Omega,5 } & =& -\gamma^T_{\Omega,5} \label{gpa} \end{eqnarray} for $\Omega $ acting on 9 and on 5-branes. The first condition is the usual requirement of global consistency of the ten form potential in Type I theory. Second equation is in agreement with the Gimon and Polchinski action, analyzed in \cite{gp}. These constraints lead to $SO(2N_9)$ and $USp(2N_5)$ groups in the 99 and 55 open string sectors respectively, where $2N_9$ $(2N_5)$ is the number of D9(5)-branes (an even number is required by $\Omega $ action we will consider). When $N_9=N_5=16$ such conditions ensure cancellation of untwisted tadpoles. Notice that consistency under the action of ${(\Omega\theta ^k)}^2 =\theta ^{2k}$ and (\ref{gpa}) lead to \begin{equation} \gamma_{k,p}^* = \gamma_{\Omega,p} \, \gamma_{k,p} \gamma_{\Omega,p} \label{famp} \end{equation} for $p=9,5$. Thus, for a $Z_N$ orbifold twist action, with $N=2P$ ($N=2P+1$) for $N$ even (odd), a generic matrix $\gamma _{\theta,p}$ can be written as \begin{equation} \gamma_{1,p}=({\tilde \gamma_{1,p}},{\tilde \gamma _{1,p}}^{*}) \end{equation} where $* $ denotes complex conjugation. ${\tilde \gamma }$ is a $N_p\times N_p$ diagonal matrix given by \begin{equation} {\tilde \gamma }_{1,p} = {\rm diag \,} (\cdots,\alpha^{NV_j}I_{n_j^p},\cdots, \alpha^{NV_ P} I_{n_P^p}) \label{gp} \end{equation} with $\alpha = {\rm e}^{2i\pi /N}$. $V_j=\frac{j}N$ with $j=0,\dots, P$ corresponds to an action ``with vector structure '' ($\gamma ^N=1$) while $V_j= \frac{2j-1}{2N}$ with $j=1,\dots,P$ describes an action ``without vector structure' ($\gamma_{1,p} ^N=-1$) \footnote{Following the classification introduced in \cite{blpssw} for six-dimensional models.}. If we choose matrices $\gamma_{\Omega,9}$ and $\gamma_{\Omega,5}$ \begin{equation} \gamma_{\Omega,9} = \left(\begin{array}{cc}0&{\rm I}_{N_9}\\ {\rm I}_{N_9} & 0 \end{array}\right) \quad ; \quad \gamma_{\Omega,5} = \left(\begin{array}{cc}0&-i{\rm I}_{N_5}\\ i{\rm I}_{N_5} & 0 \end{array}\right) \label{goms95} \end{equation} then (\ref{gpa}) and (\ref{famp}) are satisfied. In what follows we will be mainly concerned with actions ``without vector structure '' whenever D5-branes are present \footnote{This is the case most widely studied for compact orientifolds although models with vector structure can also be constructed \cite{lykken,afiv}.}. In those cases, the Chan-Paton matrices for the orbifold twist break the symplectic factors down to unitary groups. For later convenience note that the trace of twist matrix above, or in general of its $k$-th power $\gamma _{k,p}$, reads \begin{equation} {\rm Tr \,}\gamma_{k,p}=\sum_{j=0(1)}^{ P}{2n_j^p \cos(2\pi V_j k)} \label{trg} \end{equation} where the sum starts from 0(1) for the ``with (without) vector structure'' actions. In particular, for $k=0$ we obtain $Tr I_p=\sum {2n_j^p}=2N_p$, the number of Dp-branes. Moreover, in order to compare anomaly cancellation conditions, usually given in terms of the integers $n_j^p$, with tadpole equations, usually written in terms of above traces, it is also useful to have the former expressed in terms of the latter. This is easily achieved by performing a discrete Fourier transformation. For instance, for the ``without vector structure'' case, by multiplying both sides of equation (\ref{trg}) by $\cos(2k\pi V_j)$ and by summing over $k=0,\dots,N-1$ and using that ${\rm Tr \,} \gamma_{N-k,p}=-{\rm Tr \,} \gamma_{k,p}$, we obtain \begin{equation} n_j^p= \frac{1}N[ {\rm Tr \,}\gamma _{0,p} + 2\sum_{k=1}^{ P} {\rm Tr \,}\gamma _{k,p} \cos({2V_jk\pi })] \label{njp} \end{equation} for $j=1,\dots,P$. A similar expression is valid for the shift "with vector structure" (N odd) where we also have the $j=0$ term \begin{equation} 2n_0^p= \frac{1}N[ Tr\gamma _{0,p} + 2\sum_{k=1}^{ P} Tr\gamma _{k,p}] \label{nop} \end{equation} \medskip The spectrum associated to the 9 and 5-brane orientifold configuration is easily obtained by working in a Cartan-Weyl basis (see \cite{afiv}). The gauge fields living on the world-volume of a D$p$-brane have associated Chan-Paton factors $\lambda ^p$ corresponding to the gauge group $G_p$ with $G_9=SO(2N_9)$ and $G_5=Sp(2N_5)$. In Cartan-Weyl basis such generators are organized into charged generators $\lambda_a = E_a$, $a=1,\cdots, {\rm dim}\, G_p - {\rm rank}\, G_p$, and Cartan algebra generators $\lambda_I = H_I$, $I=1,\cdots, {\rm rank}\, G_p$, such that \begin{equation} [H_I, E_a]=\rho_I^aE_a \label{cw} \end{equation} where the (${\rm rank}\, G_p$)-dimensional vector with components $\rho_I^a$ is the root associated to the generator $E_a$. The matrices $\gamma_{1,p}$ and its powers represent the action of the $Z_N$ group on Chan-Paton factors, and they correspond to elements of a discrete subgroup of the Abelian group spanned by the Cartan generators. Hence, we can write \begin{equation} \gamma _{1,p}= e^{-2i\pi V^p \cdot H } \label{Vdef} \end{equation} Thus, this equation defines a (${\rm rank}\, G_p$)-dimensional vector $V^p$ with coordinates corresponding to the $V_j$'s defined in (\ref{gp}) above. Cartan generators are represented as tensor products of $\sigma_3$ Pauli matrices. In such a description the massless states are easily found. Let us consider the case in which all 5-branes sit at the origin. In the $pp$ sector the gauge group is obtained by selecting the root vectors satisfying \begin{equation} \rho^a \cdot V^p= 0 {\rm \, mod \,} {\bf Z} \label{v9p} \end{equation} while matter states correspond to charged generators with \begin{equation} \rho^a \cdot V^p= v_i {\rm \, mod \,} {\bf Z} \label{m9p} \end{equation} Recall that root vectors for orthogonal groups are of the form $\underline {(\pm1,\pm1,0,\cdots,0)}$ where underlining indicates that all possible permutations must be considered. In the symplectic case we have to include, in addition, the long roots $\underline {(\pm2,0,\cdots,0)}$. In the $95$ sector the subset of roots of $G_9\times G_5$ of the form \begin{equation} P _{(95)}= (W_{(9)}; W_{(5)})= ({\underline {\pm 1, 0, \cdots, 0}};{ \underline {\pm 1, 0, \cdots, 0}}) \label{w95def} \end{equation} must be considered. Matter states are obtained from \begin{equation} P_{(95)} \cdot V ^{(95)}= (s_jv_j +s_kv_k) {\rm \, mod \,} {\bf Z} \label{95sh} \end{equation} with $s_j=s_k=\pm \frac1{2}$, plus (minus) sign corresponding to particles (antiparticles) and $V^{95}= (V^9; V^5)$. \bigskip \section{Tadpoles versus anomalies in $D=4$, $N=1$ orientifolds} \subsection{Odd order orientifolds} In this subsection we center on the study of $D=4$ $N=1$ $Z_N$ orientifolds, with odd $N$. These models are consistent without the introduction of D5-branes, so only D9-branes are included. The only compact odd order orientifolds correspond to $Z_3$, $Z_7$ and $Z_3\times Z_3$ with vector twists given in Table~\ref{tzn}. However, let us momentarily be more general and consider also $Z_N$ orbifold actions which are not necessarily crystallographic. Thus, we focus on twist generators $\theta$ with eigenvalues $\frac{1}N(t_1,t_2,t_3)$ with $\sum_{a=1}^{3}{t_{a}}=0$ $ {\rm mod} \ N$, and $N=2P+1$. The strategy to study the relation between the anomaly and tadpole cancellation conditions will be as follows. First we compute the gauge group and massless matter for a general such model. In this step only algebraic consistency conditions (group law) are imposed on the Chan-Paton matrices. Next we find generic conditions for cancellation of gauge anomalies \footnote{Notice that when one of the gauge factors is absent, the conditions of anomaly cancellation may be less restrictive than those we consider \cite{lr}. However, this case is rather particular, and we find it is more insightful to consider `generic' cancellation of anomalies, as we do in the present paper.}. Since for $SU(n)$ groups only fundamental ${\bf n}$ and/or antisymmetric ${\bf a}_n$ representations (or their conjugates) appear, such conditions will manifest as restriction on the group ranks in order to ensure that only anomaly-free combinations are allowed. Finally, these restrictions are compared with tadpole cancellation equations obtained from type IIB orientifolds. In agreement with eq.(\ref{gp}), the generic action of the orbifold twist on 9-branes can be encoded in the matrix \begin{equation} \gamma_{1,9}=({\tilde \gamma_{1,9}},{\tilde \gamma}^{*}_{1,9}) \end{equation} with ${\tilde \gamma }$ given by \begin{equation} {\tilde \gamma } = {\rm diag \,} (I_{n_0^9}, \alpha I_{n_1^9}, \cdots, \alpha^{j} I_{n_j^9},\cdots, \alpha^P I_{n_P^9}) \label{g9} \end{equation} and $\alpha = {\rm e}^{2i\pi /N}$. The associated shift is \begin{equation} V^{9 }={\frac{1}{N}} (0,\cdots,0,1\cdots 1,\cdots,j,\cdots,j,\cdots, P, \cdots,P) \label{v9} \end{equation} from where we can easily read the gauge group to be \begin{equation} SO(u_0)\times \prod _{j=1}^P U(u_j) \label{gg9} \end{equation} where we have defined $u_0=2n_0^9$ and $u_j=n_j^9$. For the sake of clarity let us first consider orbifold twists of the form $\frac{1}N(1,1,-2)$. The corresponding massless spectrum is \begin{eqnarray} & & 2[{\overline {\bf a}}_P + ({\bf u}_0,{\bf u}_1)_{(1)} +\sum _{j=1}^ {P-1} ({\overline {\bf u}_j},{\bf u}_{j+1})_{(-1,1)}]+ \nonumber \\ & & {\overline {\bf a}}_1 +({\bf u}_0,{\overline {\bf u}}_2 )_{(-1)} + \sum _{j=1}^ {P-2}({ {\bf u}_j},{\overline {\bf u}}_{j+2})_{(-1,1)} + ({\bf u}_{P-1}, {\bf u}_P)_{(1,1)} \label{oddspec} \end{eqnarray} where, inside brackets we have indicated the charge with respect to the $U(1)$ factor in $U(u_j)$. The absence of $SU(u_j)$ gauge anomalies requires \begin{eqnarray} & & SU(u_1) \quad : 2 u_0 +u_3-2u_2-u_1+4=0 \nonumber \\ & & SU(u_j) \quad : 2u_{j-1}+u_{j+2}-2u_{j+1}- u_{j-2} =0\\ & & SU(u_P) \quad : 3u_{P-1} -u_{P -2}-2u_P+8=0 \nonumber \label{oddgac1} \end{eqnarray} where $j\ne 1, P$. Actually, by performing the identifications $u_j=u_{-j}$ and $u_{N+k} =u_k$ these conditions can be recast into the unique expression \begin{equation} SU(u_j) \quad : 2u_{j-1}+u_{j+2}-2u_{j+1}- u_{j-2} = -4\delta _{j,1}-8{\delta _{j,P}} \end{equation} for $j=1, P$. Moreover, it is not difficult to generalize these equations to the case of general odd order orbifolds, generated by a twist with eigenvalues $\frac{1}N (t_1,t_2,t_3)$, with $\sum t_{a}= 0$ ${\rm mod}$ $ N$. We obtain \begin{equation} SU(u_j) \quad : \sum _{a=1}^{3} (u_{j-t_a}-u_{j+t_a }) = 4\sum_{a=1}^{3} (\delta _{2j,t_a}-\delta _{2j,-t_a}) \label{gaco} \end{equation} for $j=1,\cdots, P$, and the arguments of the Kronecker deltas are defined mod N. Such anomaly cancellation conditions can be reexpressed in terms of traces of the matrix $\gamma$ by using equations (\ref{njp}) and (\ref{nop}) that, with the above definition of $u_0$ now read \begin{equation} u_j= \frac{1}N[ {\rm Tr \,}\gamma _{0,9} + 2\sum_{k=1}^{ P} {\rm Tr \,}\gamma _{k,9} \cos(\frac{2kj \pi} N)] \label{uj} \end{equation} for $j=0, \dots,P$ Hence, by replacing in (\ref{gaco}) we obtain \begin{equation} \frac{1}N \sum_{k=1}^{ P } \sin( \frac{2\pi kj}N)[\sum _{a=1}^3 \sin(\frac{2\pi kt_a}N ) ] Tr \gamma _{k,9} = \sum _{a=1}^{3} (\delta_{2j,t_a}-\delta _{2j,-t_a}) \label{antrgo} \end{equation} By Fourier transforming the delta functions and using that \begin{equation} \sum _{a=1}^3 \sin(\frac{2\pi k t_a }N) = -4 \prod _{a=1}^3 \sin(\frac{\pi kt_a}N ) \label{ck} \end{equation} we obtain the general conditions for the absence of gauge anomalies \begin{equation} \prod _{a=1}^3 \sin( \frac{\pi kt_a}N)\,[\, 2{\rm Tr \,} \gamma _{2k,9} \prod_{a=1}^3 \cos(\frac{\pi kt_a}M) -1 \, ] =0 \label{antrgos} \end{equation} Notice that even if $Tr\gamma _{0,9}$ appears in (\ref{uj}), it is not present in these conditions, there is no dependence on the total number of 9-branes. This was expected, given the relation of these gauge theories to systems of D3 branes at non-compact singularities, to be discussed below. For the compact $Z_3$ and $Z_7$ cases these are the well known twisted tadpole cancellation conditions $Tr \gamma _{1,9}=-4$ and $Tr \gamma _{2}=4$ respectively \cite{ang,kak3,afiv} . Such conditions on the traces or, equivalently, equations (\ref{gaco}), completely fix the values of $u_j$'s, and thus lead to the unique solutions found in the literature, when the number of 9-branes is 32. In order to interpret several results in this paper, it will be useful to relate our models to systems of D3 branes at orientifold singularities \cite{lykken,kak1,kak2,iru}. The basic observation is that the gauge theory we have described in the case at hand can be realized by placing D3 branes at a non-compact $Z_N$ orientifold singularity \cite{iru}. This can be understood by T-dualizing along the three compact directions, which transforms the D9-branes into D3-branes sitting at one of the fixed points, and then taking a decompactification limit. Thus, the conditions (\ref{antrgos}) ensure the cancellation of anomalies on the D3 brane world-volume. This explains why ${\rm Tr \,} \gamma_0$ (which in this case is the total number of D3 branes) is unconstrained: there is an infinite family of non-anomalous theories, parametrized by the number of D3 branes at the singularity. Actually, in the non-compact case, the conditions above are exactly equivalent to the tadpole cancellation conditions, in analogy with the result in \cite{lr}. We have just seen that in the particular case of $Z_N$ ($N$ odd) orientifolds, compact models also have this property. \medskip The same conclusion, however, does not follow for other types of orientifold, as we show below. To this end, let us consider for instance orientifolds $Z_{N_1}\times Z_{N_2}$, with $N_1$, $N_2$ odd. As in the above models, these theories are consistent without the addition of D5-branes. The analysis of the relation between anomaly cancellation and tadpole cancellation conditions can be studied following the strategy used above, so we will be more sketchy, leaving the details for specific examples. The first step is to compute the spectrum on the D9-brane sector for a general $Z_{N_1}\times Z_{N_2}$ model, and compute its non-Abelian anomalies in terms of the ranks of the group factors. Then we perform a discrete Fourier transform to rewrite them in terms of Chan-Paton matrices associated to the twists ${1\over N} (t_1,t_2,t_3)$ in the orientifold group. The resulting anomaly cancellation conditions have exactly the form (\ref{antrgos}) (It must be understood that the matrix $\gamma$ in (\ref{antrgos}) will correspond to a product of powers of $\gamma _{\theta}$ and $\gamma _{\omega}$ associated to the particular twist ${1 \over N}(t_1,t_2,t_3)$, where $\theta $ and $\omega$ are the $Z_{N_1}$, $Z_{N_2}$ generators, respectively). The fact that we obtain the same expression for $Z_N$ and $Z_{N_1}\times Z_{N_2}$ orientifolds (with odd $N$, $N_1$, $N_2$) is related to the fact that the twists in a $Z_{N_1}\times Z_{N_2}$ orientifold and in $Z_N$ orientifolds have the same structure (namely, no order two twist is contained in the group). In particular the expression (\ref{antrgos}) holds for the compact $Z_3\times Z_3$ orientifold with $\theta $ and $\omega$ described by the eigenvalues $\frac 13(1,-1,0)$ and $\frac 13 (0,1,-1)$. It is important to observe that the first coefficient in (\ref{antrgos}) vanishes when one direction is not affected by the twist. Therefore, for the case under consideration, only one constraint corresponding to the twist $\theta\omega ^2$ with eigenvalues $\frac 13 (1,1,-2)$, thus affecting all three complex directions, is found. It reads \begin{equation} Tr \gamma _{\theta } \gamma _{\omega ^2} =-4 \label{z3z3ac} \end{equation} Hence, we find that anomaly cancellation is much less restrictive, in this case, than tadpole cancellation, which also requires \cite{kak3,zwart} the set of equations \begin{eqnarray} {\rm Tr \,} \gamma_{\theta} &=& {\rm Tr \,} \gamma_{\omega} \ \, = \ \, {\rm Tr \,} \gamma_{\theta \omega} \ \, = \ \, 8 \label{otras} \end{eqnarray} associated to the other twists, to be satisfied. Imposing these additional conditions, the spectrum of the model is completely fixed. Thus, even if a whole set of $Z_3\times Z_3$ anomaly free models satisfying (\ref{z3z3ac}) can be constructed, there is, however, a unique solution satisfying all tadpole cancellation conditions. Let us consider this case in more detail. Twists $\theta $ and $\omega$ are represented by the matrices \begin{eqnarray} {\tilde \gamma} _{\theta } & =& {\rm diag \,} ({\rm I}_{w_0},{\rm I}_{u_1}, \alpha {\rm I}_{u_2},\alpha {\rm I}_{u_3},\alpha {\rm I}_{u_4}) \nonumber \\[0.2ex] {\tilde \gamma} _{\omega } & =& {\rm diag \,} ({\rm I}_{w_0},\alpha {\rm I}_{u_1}, {\rm I}_{u_2},{\rm I}_{u_3}, \alpha ^2 {\rm I}_{u_4}) \end{eqnarray} with $\alpha = {\rm e}^{2i\pi/3}$. The gauge group is $SO(2w_0)\times U(u_1) \dots U(u_4)$ and the massless spectrum can be easily computed to be \begin{eqnarray} & & {\overline {\bf a}}_2 + ({\overline {\bf u}_1},{\bf u}_3)+ ( {\bf u}_1,{\bf u}_4)+( {\overline {\bf u}}_3,{\overline {\bf u}}_4)+ ({\bf 2w}_0,{\bf u}_2)\nonumber \\ & & {{\bf a}}_4 + ({\overline {\bf u}_1}, {\overline {\bf u}_3})+ ( {\bf u}_1,{\overline {\bf u}}_2)+( { {\bf u}}_2,{{\bf u}}_3)+ ({\bf 2w}_0,{\overline {\bf u}}_4)\\ & & {\bf a}_1 + ({\overline {\bf u}_2},{\bf u}_4)+ ( {\bf u}_3,{\overline {\bf u}} _4)+( {\bf u}_2,{\overline {\bf u}}_3)+ ({\bf 2w}_0,{\overline {\bf u}}_1) \nonumber \label{z3z3spec} \end{eqnarray} If we only impose the anomaly cancellation condition \begin{equation} {\rm Tr \,} \gamma _{\theta} \gamma _{\omega ^2} =2(w_0+u_3)-u_1-u_2-u_4=-4 \label{z3z3acn} \end{equation} many anomaly-free spectra can be obtained. Indeed, as discussed above, these models can be related to systems of D3-branes at $Z_{3}\times Z_{3}$ singularities. For these {\it non-compact} constructions, all these models are consistent. In the non-compact case the reason why the tadpole conditions (\ref{otras}) need not be imposed is clear \cite{lr}. Those conditions correspond to twists which leave one complex plane unrotated. Thus the RR flux can escape to infinity through those planes and one does not need to impose the corresponding tadpole cancellation. In the compact case this is different: even though those planes are unrotated they are compact and the RR charge is trapped. Thus the twisted RR charge corresponding to those twists has to be canceled and the conditions (\ref{otras}) have to be imposed. However, although these conditions are needed for consistency, they are totally disconnected from the issue of gauge anomaly cancellation. \subsection{Even order orientifolds} For orientifolds including even order twists, there will be in general both D9-branes and D5-branes present. In this section we consider the construction of $D=4$, $N=1$ models obtained from orientifold configurations of D9 and D5-branes. As in the previous section, we first consider a general (not necessarily crystallographic) orbifold twist. In a first step we compute the gauge group and massless spectrum of such models for an arbitrary number of $2N_9$ ($2N_5$) D9(D5)-branes. Generic conditions for cancellation of gauge anomalies are then found. Using the discrete Fourier transform, these restrictions are then compared with tadpole cancellation equations obtained from type IIB orientifolds. Finally, explicit solutions to these equations are discussed. In some cases new solutions, other than those reported in the literature are found. For simplicity we concentrate on models with only one set of D5-branes, all of them sitting at the fixed point at the origin. Other situations are considered through specific examples. Let us consider arbitrary $Z_N$ ($N=2P$) twists with eigenvalues given by $\frac{1}N (t_1,t_2,t_3)$ with $t_1+t_2+t_3=0$ and $t_3$ an even integer (thus $t_1,t_2$ are odd). We concentrate on models without vector structure. Thus, from eq. (\ref{gp}) we have \begin{equation} {\tilde \gamma } = {\rm diag \,} (\alpha I_{n_1},\cdots,\alpha^{(2j-1)}I_{n_j},\cdots, \alpha^{(2P-1)} I_{n_P}) \label{g5} \end{equation} with $\alpha = {\rm e}^{i\pi /N}$ and where we have dropped the upperindex 5 ($n_j^5=n_j$). A similar structure is used in the 9-brane sector just by replacing the integers $n_j$'s by $u_j$'s. These matrices correspond to the shifts \begin{equation} V^{p }={\frac{1}{2N}} (1,\cdots 1,\cdots,2j-1,\cdots,2j-1,\cdots,2P-1, \cdots, 2P-1) \label{vp} \end{equation} with $n_j (u_j)$ entries $(2j-1)$ for $p=5 (9)$, with $j=1,\cdots,P$. The resulting gauge group is \begin{equation} \prod _{j=1}^{P} U(n_j)\times \prod _{j=1}^{P} U(u_j) \label{gg5} \end{equation} The massless states in $55$ sector are \begin{eqnarray} \sum _{a=1}^3 {\sum_{j=1}^{P}}^{\prime} & & [{({\overline {\bf n}_j},{\bf n}_{j+t_a})_{(-1,1)}}+ {({\overline {\bf n}_j},{\overline {\bf n}}_{[-j-t_a+1]})_{(-1,-1)}}+\nonumber \\ & & ({\bf n}_j,{\overline {\bf n}}_{j-t_a})_{(1,-1)}+ ({\bf n}_j,{ {\bf n}}_{[-j+t_a+1]})_{(1,1)}] +\nonumber \\ {\sum _{j=1}^{P}} & & [{{\bf a}_j}_{(2)} (\delta _{j,{\frac{t_1+1}2}}+ \delta _{j,{\frac{N+t_1+1}2}})+ {{\overline {\bf a}}_j}_{(-2)} (\delta _{j,{\frac{-t_1+1}2 }}+ \delta _{j,{\frac{N-t_1+1}2} })+ \nonumber\\ & & (t_1 \rightarrow t_2 )] \label{z55} \end{eqnarray} while the $59$ sector spectrum is \begin{eqnarray} & & \sum _{j=1}^ {P}{({\overline {\bf n}_j},{\bf u}_{j+t})_{(-1,1)}}+ {({\overline {\bf n}_j},{\overline {\bf u}}_{[-j-t+1]})_{(-1,-1)}}+ \nonumber \\ & & {({\bf n}_j,{\overline {\bf u}}_{j-t})_{(1,-1)}}+ {({\bf n}_j,{ {\bf u}}_{[-j+t+1]})_{(1,1)}}+ ({\bf n} \leftrightarrow {\bf u}) \label{specfn} \end{eqnarray} where the indices $j$, $j^{\prime}$ are defined mod $N$. Notice that sums over bifundamentals with indices $(j,j')$ must be performed such that $1\le j\le j'\le \frac{N}2$. Also, the primed sum on 55 sector means that $j=j'$ is not included. In fact, in this case, antisymmetric representations do appear for $t_a$ odd, i.e. $a=1,2$ for our convention above. In 59 sector we have defined $t=\frac{t_1+t_2}2$, see (\ref{95sh}). The spectrum in 99-sector can be obtained from eq.(\ref{z55}) by replacing $n$'s by $u$'s. Again, it is straightforward to write down a general condition for the cancellation of non-Abelian anomalies. It reads \begin{eqnarray} SU(n_j) \quad & : & \sum _{a=1}^3 [(n_{j-t_a}-n_{j+t_a})+u_{j-t}-u_{j+t} ]= 4 [\delta _{j,{\frac{t_1+1}2}}+ \delta _{j,{\frac{N+t_1+1}2}} \nonumber\\ & & -\delta _{j,{\frac{-t_1+1}2 }}- \delta _{j,{\frac{N-t_1+1}2} }+ (t_1 \rightarrow t_2) ] \label{gacg} \end{eqnarray} with $j=1,\dots,{P}$ and where the identifications $n_{j}=n_{-j+1}=n_{j+N}$ are understood. As in the odd twist case, such equivalences are automatically implemented when $n_j$'s ($u_j$'s) are expressed in terms of traces through ec.(\ref{njp}). In fact, by using such equation above we find that, in order to ensure the absence of gauge anomalies for the 55 group sector we must have \begin{equation} \frac{1}N \sum_{k=1}^{ P-1 }\sin (2\pi k V_j)[\sin(\frac{t_3\pi k}N ) A_k]= -\delta _{j,{\frac{t_1+1}2}}-\delta _{j,{\frac{N+t_1+1}2}} +\delta _{j,{\frac{-t_1+1}2 }}+ \delta _{j,{\frac{N-t_1+1}2} } + (t_1 \rightarrow t_2) \label{antrg} \end{equation} with \begin{equation} A_k= 4\sin(\frac{t_1\pi k}N) \sin(\frac{t_2\pi k}N ) {\rm Tr \,}\gamma _{k,5,0} + {\rm Tr \,}\gamma _{k,9} \label{akg} \end{equation} where we have emphasized, by adding a $0$ subscript, that D5-branes are located at the origin. Again $\delta$ functions appear from the contribution of antisymmetric representations. Notice that the sum is actually up to $P-1$ (and not up to $P$). This is due to the fact that, from our definition (\ref{g5}) of the twist matrix , in the case without vector structure , \begin{equation} Tr{\gamma_{P,p}}=0 \label{o2t} \end{equation} This equation for the order two twist matrix, which in this case is automatically satisfied by construction, appears in \cite{afiv} (eq.(2.39)) as necessary condition for the cancellation of tadpoles Again, it will be useful to keep in mind that these gauge theories can be realized on the world-volume of D3-branes at $Z_N$ orientifold singularities. In this case, D7-branes are also present in the configuration. In this non-compact context, it is possible to show that the anomaly cancellation conditions (\ref{antrg}) are exactly equivalent to the tadpole cancellation conditions. In the following we show that this property is in general no longer true in the compact models. The constraints that we have just obtained may be read as a set of $P$ equations (for each value of $j$) with $P-1$ unknowns $\sin(\frac{t_3\pi k}N ) A_k$. Thus, the system is, in principle, overdetermined and it could have no solutions at all, unless not all equations are really independent. In fact, we will see that some models are not consistent. Recall that a similar set of equations, obtained from above simply by exchanging $\gamma _{k,9}$ and $\gamma _{k,5}$ would be required in order to avoid gauge anomalies in the 99 sector groups. This $5 \leftrightarrow 9$ symmetry is due to the fact that all fivebranes have been put at the origin, and it is a manifestation of T (self) duality. Nevertheless we will see that despite the symmetry of anomaly equations, in some cases they allow for solutions which are not symmetric under the exchange $5 \leftrightarrow 9$. The additional tadpole cancellation equation leads to a fully T duality invariant spectrum. The above constraints are valid for an arbitrary, not necessarily chrystalographic, even twist action on 5 and 9-branes. In what follows we analyze the compact cases shown in table 1. The global constraint $Tr\gamma _{0,9}=Tr\gamma _{0,5}= 32$ must be imposed in such cases. \bigskip \bigskip {\bf i) Non consistent models} For orbifold groups containing a twist with eigenvalues $\frac{1}4(1,1,-2)$, namely the orbifold actions { $Z_4$, $Z_8$, $Z_8$'} and $Z_{12}$' the above equations have no solutions. In fact, these orientifolds were found in \cite{afiv} to be ill-defined. Difficulties stem from the presence of a Klein-bottle tadpole proportional to the volume $V_3$ of the third compact dimension. Interestingly enough, such inconsistencies are recovered here from the point of view of anomaly cancellation. This can be easily seen in the $Z_4$ case. In fact, (\ref{antrg}) leads to the incompatible equations $\frac{\sqrt2}2 A_1=8$ for $j=1$ and $\frac{\sqrt2}2 A_1=-8$ for $j=2$ where $A_1= 2Tr\gamma _{1,5,0} + Tr\gamma _{1,9}$. The same situation repeats in the other cases. We should emphasize that the above comments for inconsistency refer only to Chan-Paton twists without vector structure. It is in principle possible to find consistent $D=4$, $N=1$ orientifolds corresponding to these twists but with CP twists {\it with } vector structure. Indeed we have found consistent (non-chiral) examples based on $Z_4$ with vector structure. \bigskip \bigskip {\bf ii) Consistent models} For all other twists in Table 1 it is possible to write down a general, consistent, solution to eq. (\ref{antrg}). Namely \begin{equation} \sin(\frac{t_3\pi k}N )A_k= -4(1+(-1)^k )[\sin(\frac{t_1\pi k}N) + \sin(\frac{t_2\pi k}N )] \label{aksol} \end{equation} with $k=1,\dots P$. Thus, for each twist $\theta ^k$ of order $k$ we obtain a condition which relates ${\rm Tr \,}\gamma _{k,5}$ and ${\rm Tr \,}\gamma _{k,9}$ as they show up in tadpole cancellation equations. Whenever $k$ is such that $\theta ^k $ twist leaves the direction parallel to the D5-brane untouched, then $\sin(\frac{t_3\pi k}N )=0$ (thus also $\sin(\frac{t_1\pi k}N) + \sin(\frac{t_2\pi k}N) =0$) and no constraint is present for the corresponding ${\rm Tr \,}\gamma _{k,9}$ and ${\rm Tr \,}\gamma _{k,5}$. On the other hand, since we must have a similar solution for anomaly cancellation in $99$ sector groups but with $5\rightarrow 9$ in (\ref{akg}), we will have \begin{equation} [ 4\sin(\frac{t_1\pi k}N) \sin(\frac{t_2\pi k}N )-1] ({\rm Tr \,}\gamma _{k,5,0} - {\rm Tr \,}\gamma _{k,9})=0 \label{t9t5f} \end{equation} for all other twists with $t_3\pi k \ne 0$ $mod$ $N$. Notice that $4\sin(\frac{t_1\pi k}N )\sin(\frac{t_2\pi k}N )$ is, by the Lefschetz fixed point theorem, nothing but the number of fixed points of twist $\theta ^k$ in the complex directions $(Y_1,Y_2)$. Thus, this equation implies that whenever points other than the origin are kept fixed then ${\rm Tr \,}\gamma _{k,5,0} = {\rm Tr \,}\gamma _{k,9}$. Also notice that only even values of $k$ could lead to non zero values for $A_k$. Let us consider the different cases in more detail: \bigskip \bigskip ${\bf Z_6}$: From eq.(\ref{aksol}) we find $A_1=0$ and $A_2=16$, namely \begin{eqnarray} & & {\rm Tr \,} \gamma_{1,9} + {\rm Tr \,} \gamma_{1,5,0} = 0 \nonumber\\ & & {\rm Tr \,} \gamma_{2,9} + 3{\rm Tr \,} \gamma_{2,5,0} \ \ = \ \ 16 \label{z6ac} \end{eqnarray} These are tadpole cancellation conditions found in \cite{kak3,afiv}. However, in reference \cite{afiv} it is found that tadpole cancellation at other fixed points, other than the origin, imposes an extra requirement. Namely \begin{equation} {\rm Tr \,} \gamma_{2,9} + 3{\rm Tr \,} \gamma_{2,5,J} \ \, = \ \, 4 \label{tcz61} \end{equation} with $J=1,\cdots, 8$ denoting the fixed points of $\theta^2$ in the $(Y_1,Y_2)$ planes, other than the origin, where 5-branes may sit. Since here all branes are at the origin then ${\rm Tr \,} \gamma_{2,5,J}=0$. We are led to \begin{equation} {\rm Tr \,} \gamma_{2,9} = \ \, 4 \label{tcz62} \end{equation} This is precisely the extra condition (\ref{t9t5f}) that results when anomaly cancellation for the $99$ sector groups is required. Thus, we learn that for the $Z_6$ orientifold, generic absence of anomalies and of tadpole divergences are equivalent. Let us stress that to arrive at this conclusion, analysis of tadpole cancellation at all fixed points (and not only at the origin) is needed. In fact, it is equation (\ref{tcz62}) which makes tadpole equations to look symmetric under the exchange of 9 and 5-branes (all at the origin) and thus ensures cancellation of 99 sector anomalies. It appears to us that this point is not sufficiently clear in the literature. Sometimes tadpole equations asymmetric between D9 and D5-branes are written down for four- or six-dimensional orientifolds. These are misleading since, as they stand, they would allow for 5-9 asymmetric solutions which could be anomalous. Taking into account tadpole cancellations in fixed points away from the origin does in fact impose D9-D5 symmetric solutions. This is just the right behaviour, given the fact that putting all 5-branes at the origin is a selfdual configuration under T-duality. \medskip Let us now look for solutions of the above equations. An interesting feature is that the first row in (\ref{z6ac}) indicates the possibility of a local cancellation among 9-branes and 5-branes RR charges that may not vanish independently. This possibility has not been noticed in the literature, and we will exploit it here to construct new solutions of the tadpole equations for all 5-branes at origin. The new possibilities, as we saw in eq.(\ref{t9t5f}), have to do with the fact that only the origin is fixed under $Z_6$ twist. In order to study the solutions it is easier to rewrite the traces in terms of $n_j$'s and $u_j$'s. Hence, by using (\ref{trg}) we have \begin{eqnarray} \frac{1}{\sqrt3} {\rm Tr \,} \gamma_{1,5} & = & n_1-n_3 \nonumber\\ {\rm Tr \,} \gamma_{2,5} & = & n_1-2 n_2+n_3 \end{eqnarray} and similar equations for ${\rm Tr \,} \gamma_{k,9} $ in terms of $u_j$'s. Replacing (\ref{z6ac})-(\ref{tcz62}) and using that there are 32 D5- and D9-branes we obtain \begin{eqnarray} u_1& = & 12-n_1\nonumber \\ u_2& = & n_2= 4 \nonumber \\ u_3& =& n_1 \label{a9} \end{eqnarray} Therefore, from (\ref{gg5}), the gauge groups, depending on the free integer parameter $n_1 \le 12$, are \begin{equation} [U(n_1)\times U(4)\times U(12-n_1)]_{55} \times [U(12-n_1)\times U(4)\times U(n_1) ]_{99} \label{ggz6} \end{equation} The matter content may easily be obtained from (\ref{z55}) and (\ref{specfn}) . Notice that the solution given in \cite{kak3,afiv} has $n_1=n_3=6$ implying that ${\rm Tr \,} \gamma_{1,5}={\rm Tr \,} \gamma_{1,9}=0 $. Therefore it corresponds to the case in which 5 and 9 brane contributions cancel independently. \bigskip ${\bf Z_6}${\bf '}: Absence of anomalies in the 55 sector leads to \begin{eqnarray} & & {\rm Tr \,} \gamma_{1,9} - 2{\rm Tr \,} \gamma_{1,5,0} = 0 \nonumber\\ & & {\rm Tr \,} \gamma_{2,9} = -8 \label{z6'1} \end{eqnarray} which again coincide with equations for tadpole cancellation at the origin. When the anomaly cancellation conditions from the 99 sector are included the full set of constraints \begin{eqnarray} & & {\rm Tr \,} \gamma_{1,9} ={\rm Tr \,} \gamma_{1,5} = 0 \nonumber\\ & & {\rm Tr \,} \gamma_{2,9} = {\rm Tr \,} \gamma_{2,5} = -8 \label{z6'2} \end{eqnarray} is obtained. It is completely equivalent to cancellation of all twisted tadpoles, including those at the other three fixed points in the transverse directions. One aspect of this model may appear puzzling. Since the order three twist leaves one complex plane unrotated, one would expect, as it happened in the $Z_3\times Z_3$ case, that the tadpole constraint for $\gamma _{2,9}$ is not required for anomaly cancellation. It turns out that the presence of D5-branes makes things different. Indeed, cancellation of gauge anomalies for 99 sector groups requires eq.(\ref{aksol}) but with 9 and 5 indices exchanged in (\ref{akg}). Explicitly, \begin{equation} A_k= 4\sin(\frac{\pi k}6)\sin(\frac{-3\pi k}6 ) {\rm Tr \,}\gamma _{k,9} + {\rm Tr \,}\gamma _{k,5} \label{akz'6} \end{equation} and thus, in effect, no constraint on ${\rm Tr \,}\gamma _{k,9}$ is found for $k=2$. Nevertheless ${\rm Tr \,}\gamma _{2,5}=-8$ is still required. The condition on ${\rm Tr \,}\gamma _{2,9}$ is really obtained from the 55 sector consistency. A further discussion on this point is presented in section 5. Unlike $Z_6$, in this case charges among 9-branes and 5-branes must cancel independently. This is due, essentially, to the presence of extra fixed points. As in the odd orientifold cases, since now the number of conditions on the traces and the number of unknowns are the same, the solution presented in \cite{zwart,afiv} (for all 5-branes at the origin) is unique. It is instructive to analyze, from the point of view of anomaly cancellation, the situation when part of the 32 D5-branes live at other fixed points. As an example consider the case when five branes are distributed in groups of $N^5_L$ branes among the four fixed points $L=0,\dots,3$ of the $\theta $ twist in the first two planes. The structure of the $55 _L$ gauge group and matter content at point $L$ is exactly the same as that for the origin considered above. The spectrum at each point is obtained from (\ref{z55}) and (\ref{specfn}) just by replacing $n \rightarrow n^L$, with the condition $\sum_{L=0}^3 (n_1^L+n_2^L +n_3^L)=16$. Thus the anomaly constraints eq.(\ref{z6'1}) must now be imposed at each point. Namely \begin{eqnarray} & & {\rm Tr \,} \gamma_{1,9} -2{\rm Tr \,} \gamma_{1,5,L} = 0 \nonumber\\ & & {\rm Tr \,} \gamma_{2,9} = -8 \label{z6'L55} \end{eqnarray} for $L=0,\dots,3$. However, constraints involving several fixed points are obtained when the 99 group is considered. The $59$ sector (\ref{specfn}) reads \begin{equation} \sum _{L=0}^3 [ ({\overline {\bf u}}_1, {\overline {\bf n}}_1^L)+ ({ {\bf u}}_1, {\overline {\bf n}}_2^L) + ({ {\bf u}}_2, {\overline {\bf n}}_3^L) + ({{\bf u}}_3, { {\bf n}}_3^L) + ({\overline {\bf u}}_2, { {\bf n}}_1^L)] \end{equation} and so, a sum of contributions from all fixed points supporting D5-branes appears. Thus anomaly constraints will read \begin{eqnarray} & &\sum_{L=0}^3 {\rm Tr \,} \gamma_{1,5,L} -2{\rm Tr \,} \gamma_{1,9} = 0 \\ & & \sum_{L=0}^3 {\rm Tr \,} \gamma_{2,5,L} = -8 \label{z6'L99} \end{eqnarray} Eqs.(\ref{z6'L55}) and (\ref{z6'L99}) reproduce the tadpole cancellation equations for this general case. Since D5-branes are now distributed among different fixed points we expect to be able to achieve charge cancellation in different ways rather than in the unique form found above. For instance, first row in equations above tells us that now it could still be possible to achieve charge cancellation among 9 and 5-branes even if they did not vanish independently. By writing the traces in terms of group ranks (\ref{trg}) and recalling that $u_1+u_2+u_3=16$, $\sum _{L=0}^3 n_1^L+n_2^L+n_3^L=16$ we find the following consistency constraints \begin{eqnarray} & & n_1^L-n_3^L = u_1-6= 2-u_3 \nonumber \\ & & u_2 = \sum_{L=0}^3 n_2^L= 8 \nonumber \\ & & \sum_{L=0}^3 n_1^L+n_3^L = 8 \label{z'6c} \end{eqnarray} \bigskip ${\bf Z_{12}}$ In this example tadpole cancellation appears to be stronger than anomaly cancellation. Indeed, anomaly cancellation in the 55 sector gives \begin{eqnarray} A_k & = & {\rm Tr \,} \gamma_{k,9} - {\rm Tr \,} \gamma_{k,5,0} \ \, = \ \, 0 \quad ; \quad k=1,2,5 \nonumber \\[0.2ex] A_4 & = & {\rm Tr \,} \gamma_{4,9} + 3{\rm Tr \,} \gamma_{4,5,0} \ \, = \ \, 16 \label{z12t3} \end{eqnarray} which must be supplemented with the extra, independent, constraint from the equivalent equations in the 99 sector, \begin{equation} {\rm Tr \,} \gamma_{4,9} = {\rm Tr \,} \gamma_{4,5,0} \ \, = \ \, 4 \end{equation} Notice, however, that there is no constraint for $A_3={\rm Tr \,} \gamma_{3,9} + 2{\rm Tr \,} \gamma_{3,5,0} \;$, due to the appearance of the vanishing factor $\sin(\frac{4\pi k}{12})$ (for $k=3$) associated to the fourth order twist $\frac{1}4(1,-1,0) $ which leaves the third plane invariant. Thus, we expect the model will have additional tadpole conditions, not needed to ensure anomaly cancellation. This additional constraint is the tadpole condition (2.45) of \cite{afiv} \begin{equation} {\rm Tr \,} \gamma_{3,9} + 2{\rm Tr \,} \gamma_{3,5,L} = 0, \label{z12pt2} \end{equation} where $L$ refers to the fixed points of $\theta^3$ in the first and second complex coordinates. In our specific case with all D5 branes at the origin, the condition reads \begin{equation} {\rm Tr \,} \gamma_{3,9} = {\rm Tr \,} \gamma_{3,5,0} = 0 \label{tcup} \end{equation} These equations constrain the model beyond mere anomaly cancellation. This resembles much what happened with $Z_3\times Z_3$. Again, due to the fact that only the origin is fixed under $Z_{12}$ there are more solutions to the tadpole equations than those in the literature. Let us be more explicit. The generic gauge group is $\prod _{i=1}^6 U(n_i)$ while the 55+59 massless spectra is given by (\ref{z55}-\ref{specfn}) \begin{eqnarray} {\bf a}_1 + {\overline {\bf a}_6} +({\overline {\bf n}_1},{\bf n}_2)+ ({\overline {\bf n}_2},{\bf n}_3)+ ({\overline {\bf n}_3},{\bf n}_4)+({\overline {\bf n}_4},{\bf n}_5)+({\overline {\bf n}_5},{\bf n}_6) \nonumber\\ {\overline {\bf a}_3} + {\bf a}_4+ ({\overline {\bf n}_1},\overline {\bf n}_5)+({\overline {\bf n}_2},\overline {\bf n}_4) +({ {\bf n}_2},{\bf n}_6)+( {\bf n}_3,{\bf n}_5) +( {\bf n}_1;\overline {\bf n}_6)+ \nonumber \\ ({\bf n}_1,\overline {\bf n}_4)+({\bf n}_2,{\bf n}_3)+({ \overline {\bf n}_1},{\bf n}_5)+ ({ \overline {\bf n}_2},{\bf n}_6)+ ({ \overline {\bf n}_3},{ \overline {\bf n}_6}) \nonumber \\ (\overline {\bf u}_1;\overline {\bf n}_2)+ ( {\bf u}_1;\overline {\bf n}_3)+( {\bf u}_2;\overline {\bf n}_4)+({\bf u}_3;\overline {\bf n}_5)+ ({\bf u}_4;\overline {\bf n}_6)+( {\bf u}_5;\overline {\bf n}_6) +u\rightarrow n \label{z12spec} \end{eqnarray} By rewriting the traces above in terms of the group ranks we find that anomaly cancellation equations (\ref{z12t3}) constrain such ranks to satisfy \begin{eqnarray} & & n_1-n_4= u_1-u_4 \nonumber \\ & & n_1+n_3=u_1+u_3 \nonumber\\ & & n_1+n_2 = u_1+u_2 \nonumber\\ & & n_2+n_5=u_2+u_5=4\nonumber\\ & & n_1+n_3+n_4+n_6=u_1+u_3+u_4+u_6=12 \label{a8} \end{eqnarray} Thus we see that all $u_j$'s may be expressed in terms of 55 integers, apart from $u_1$, which is a free integer parameter. It is interesting to notice that if only eq. (\ref{z12t3}) are used then anomaly free (but not tadpole-free !) models, which have an asymmetric content in the $55$ and $99$ sector are possible. For instance, a family of models obeying such constraint but not (\ref{tcup}) is obtained by choosing \begin{eqnarray} n_1& = &n_4=n_5=4\\ n_2 &= & 0, \ \ n_3=k, \ \ n_6=4-k\\ u_1&=& u_4=u_5=0 \\ u_2 & = & 4, \ \ u_3=k+4, \ \ u_6=8-k \label{z12asim} \end{eqnarray} with $k\le 4$. The corresponding gauge groups are \begin{equation} [U(4) \times U(k+4) \times U(8-k)]_{99} \times [U(k) \times U(4-k) \times U(4)^3]_{55} \end{equation} These models are chiral and anomaly free, but are not symmetrical with respect to the exchange between D9 and D5 branes, and so violate (\ref{tcup}). In fact, Eq. (\ref{tcup}) reads $n_1-n_2+n_4=u_1 -u_2+u_4=4$ and thus imposes $u_j=n_j$ for all $j=1,\dots, 6$. With this extra condition the resulting gauge group is now \begin{equation} [ U(n_1) \times U(n_2) \times U(n_3) \times U(4+n_2-n_1) \times U(4-n_2) \times U(8-n_2-n_3)]_{55} \times [ same ]_{99} \label{ggz12} \end{equation} symmetric under $5\leftrightarrow 9$. These are the solutions which obey all tadpole cancellation conditions. For $n_1=n_3=3$ and $n_2=2$ one recovers the solution given in \cite{afiv}. \section{Anomalous $U(1)$'s } It is known that orientifold compactifications lead to spectra which usually contain several Abelian factors with non vanishing triangle anomalies. In \cite{iru} a generalized Green-Schwarz mechanism ensuring cancellation of such terms was found \footnote{See \cite{sagnanom,blpssw,intri} for the analogous mechanism in six dimensions.}. It involves RR scalars coming from the twisted closed sector of the string. In this section we would like to indicate how $U(1)$ anomalies actually cancel in the models we have considered above, whenever generic cancellation of non Abelian anomalies is ensured. \subsection{Mixed non-Abelian anomalies} For concreteness, let us consider the case of even orientifold models. The arguments below are also valid for odd orientifolds. In this subsection we center on the mixed anomaly of the $U(1)$ factor in $U(n_i)$ in the Dp-brane sector with $SU(n_j)$ non-Abelian group in the Dq-brane sector, denoted $T^{pq}_{ij}$. Following \cite{iru} this triangle anomaly will be canceled if $T^{pq}_{ij}+A^{pq}_{ij}=0$, where $A^{pq}_{ij}$ is the Green-Schwarz term given by \begin{equation} A_{ij}^{pq}\ = \ {1\over N}\ \ \sum_{k=1}^{N-1} \ C_k^{pq }(v)\ n_i^p\sin2\pi kV^{p }_i \ \cos2\pi kV^{p}_j \label{masterorient} \end{equation} Here $k$ runs over twisted $Z_N$ sectors, $p$, $q$ run over 5,9 (meaning 5- or 9-brane origin of the gauge boson) and \begin{eqnarray} C_k^{pp } & = & \prod _{a=1}^3 2\sin\pi kv_a \quad {\rm for}\;\; p=q \nonumber\\ C_k^{59 } & = & 2\sin\pi kv_3 \label{ckpp59} \end{eqnarray} In principle, this factorization of $U(1)$ anomalies could lead to new constraints on the spectrum of the model, beyond those imposed by generic cancellation of non-Abelian anomalies. In the following, we show this is {\em not} the case. Sum over $k$ in (\ref{masterorient}) can be performed explicitly. Consider the $p=q$ sector. By using (\ref{ck}) and orthogonality of cosines we find \begin{equation} A_{ij}^{pp}\ = -\frac {n_i}2 \sum_{a=1}^3 [\delta _{i,j+t_a}+\delta _{i,-j+t_a+1}- \delta _{i,-j-t_a+1}- \delta_{i,j-t_a}] \label{aijpp} \end{equation} and similarly \begin{equation} A_{ij}^{59} = \frac {n_i}2 [ \delta _{i,j+{\frac{ t_3}2 } }+ \delta_{i,-j+{\frac{t_3}2}+1}- \delta _{i,-j-{\frac{t_3}2}+1}- \delta_{i,j-{\frac {t_3}2}}] \label{aij59} \end{equation} where arguments of the Kronecker $\delta$ functions are defined mod $N$. It is straightforward to check from the spectrum given in (\ref{z55}) and (\ref{specfn}) that this indeed cancels the triangle anomaly whenever $i\ne j$. The $i=j$ case is a little bit more involved due to contributions from antisymmetric representations. Recall that the contribution to the triangle anomaly from the antisymmetric ${{\bf a}_j }_{(2)}$ of $SU(n)$ is, in our normalization, $2 (\frac{n-2}2)$. Thus, the $U(1)_j$-$SU(n_j)$ triangle anomaly reads \begin{eqnarray} T_{jj}^{pp}\ & = & \frac {n_j}2 \; [\; \delta _{j,{\frac{t_1+1}2}}+ \delta_{j,{\frac{N+t_1+1}2}} -\delta _{j,{\frac{-t_1+1}2 }}- \delta_{j,{\frac{N-t_1+1}2} }+ (t_1 \rightarrow t_2) \; ]+ \nonumber\\ & & \frac{1}2 \{\;\sum _{a=1}^3 (n_{j-t_a}-n_{j+t_a})+u_{j-t}-u_{j+t} + \\ & & -4 \;[\;\delta _{j,{\frac{t_1+1}2}}+ \delta _{j,{\frac{N+t_1+1}2}} -\delta_{j,{\frac{-t_1+1}2 }}- \delta _{j,{\frac{N-t_1+1}2} }+ (t_1 \rightarrow t_2)\;] \;\} \nonumber \label{tjj} \end{eqnarray} The term between curly brackets is nothing but the expression (\ref{gacg}) for cancellation of general non-Abelian anomalies. Since it must vanish the remaining contribution cancels the expression (\ref{aijpp}) for the case $i=j$. This shows that the condition of cancellation of non-Abelian anomalies implies the appropriate factorization of $U(1)$ anomalies. It is also easy to check that cubic $U(1)$ anomalies are similarly canceled by the GS mechanism without the need of further constraints. This behaviour could have been guessed, based on our results in Section~3. There we noticed that the conditions of non-Abelian anomaly cancellation are equivalent to the tadpole cancellation conditions of a system of D3-branes at non-compact orientifold singularities (possibly in the presence of D7 branes). In this language, the additional tadpole conditions only arise in compact models, where the RR flux cannot escape to infinity. Thus the additional conditions are not required for consistency of the D3 brane system. Since the cancellation of $U(1)$ anomalies is required for consistency even in the non-compact case, it cannot depend on any of the additional constraints. Thus, appropriate factorization of the $U(1)$'s must follow from the conditions already imposed by cancellation of non-Abelian anomalies, as we have shown above. \subsection{Mixed gravitational anomalies} We can use a similar computation to show that the cancellation of mixed $U(1)$-gravitational anomalies does not impose further constraints. This may appear puzzling at first sight, since gravitational anomalies are relevant only for compact models. One could expect their cancellation would involve some (or even all) of the additional tadpole cancellation conditions. So let us give an intuitive argument (of partial validity) to understand this fact before entering the detailed computation. Again, we use the realization of these gauge theories in terms of D3-branes at orientifold singularities. Certainly, in the non-compact limit gravity propagates in all ten dimensions, not just in the four-dimensions where the gauge theory lives, and so there is no obvious reason why the mixed $U(1)$-gravitational anomalies should cancel. However, some of these singularities can be embedded in compact Calabi-Yau spaces (not necessarily toroidal orbifolds). This does not change the spectrum of the field theory on the D3-branes, since they only feel local physics, but has the effect that gravitational anomalies become relevant, and must cancel. Since the mixed $U(1)$-gravitational anomaly receives contributions only from fields living on the D3 branes, it follows that gravitational anomalies must cancel for non-compact singularities if they can be embedded in a global model. Following the argument in the previous subsection, this shows that, for these singularities, factorization of gravitational anomalies must be automatic once cancellation of non-Abelian anomalies is imposed. Obviously, this argument does not apply to general singularities, since most singularities cannot be embedded in global contexts. The explicit computation below, however, shows the conclusion concerning the absence of new constraints is indeed true for any singularity. Let us consider the case of even order orientifolds (odd order orientifolds can be analyzed similarly). Starting from the spectrum given in section 3.2, the mixed anomaly $T_i^{grav.}$ for the $i^{th}$ $U(1)$ factor (in the 55 sector) can be shown to be \begin{eqnarray} T_i^{grav.} & = & \sum_{a=1}^3 (n_{i-t_a} n_i - n_i\, n_{i+t_a}) - n_i u_{i+t} + n_i u_{i-t} \nonumber \\ & & - n_i\;[\; \delta_{i,\frac{t_1+1}{2}} + \delta_{i,\frac{N+t_1+1}{2}} -\delta_{i,\frac{-t_1+1}{2}} - \delta_{i,\frac{N-t_1+1}{2}} + (t_1 \to t_2) \;] \label{angra1} \end{eqnarray} The gravitational anomaly for $U(1)$ from the 99 sector has the same structure, with the replacement $n_i \leftrightarrow u_i$. For future convenience, let us note that by using the condition for the cancellation of non-Abelian anomalies (\ref{gacg}), we can rewrite (\ref{angra1}) as \begin{eqnarray} T_i^{grav.} & = & 3n_i \;[\; \delta_{i,\frac{t_1+1}{2}} + \delta_{i,\frac{N+t_1+1}{2}} - \delta_{i,\frac{-t_1+1}{2}} - \delta_{i,\frac{N-t_1+1}{2}} + (t_1 \to t_2) \;] \label{angra2} \end{eqnarray} In \cite{iru} it was proposed that the GS contribution canceling this anomaly has the form \begin{eqnarray} {A_i^{grav}} & = & \frac 34 \frac 1N \sum_{k=1}^N \left[ C_k^{55}(v)\, {\rm Tr \,} \gamma_{k,5} + C_k^{59}(v)\, {\rm Tr \,} \gamma_{k,9} \right] n_i \sin 2\pi k V_i \label{irugrav} \end{eqnarray} This can be Fourier-transformed in a by now familiar way, and be expressed as \begin{eqnarray} {A_i^{grav}} & = & -\frac 38 n_i [\; \sum_{a=1}^3 (n_{i-t_a} + n_{-i+1+t_a} - n_{i+t_a} -n_{-i+1-t_a}) + \nonumber \\ & & (u_{i-t} + u_{-i+1+t} - u_{i+t} + u_{-i+1-t} ) \;] \end{eqnarray} Recalling the relations $n_i=n_{-i+1}$, $u_i=u_{-i+1}$, this reads \begin{eqnarray} {A_i ^{grav.}} & = & -\frac 34 n_i \left[ \;\sum_{a=1}^3 (n_{i-t_a} - n_{i+t_a}) + (u_{i-t} - u_{i+t}) \right] \end{eqnarray} which, after imposing the non-Abelian anomaly cancellation conditions (\ref{gacg}), precisely cancels (\ref{angra2}). \medskip As an explicit example of the above discussion consider the $Z_{12}$ orientifold. The GS factors eq.(\ref{irugrav}) are $A_1 ^{grav}=-3n_1$, $A_3^{grav}=3n_3$, $A_4^{grav}=-3n_4$ and $A_6^{grav}=3n_6$ while $A_2^{grav}=A_5^{grav}=0$. The mixed gravitational anomalies $T_i$ are easily computed from eq.(\ref{z12spec}). For instance for the first and second $U(1)$ factors in the 55 sector we obtain \begin{eqnarray} &T_1^{grav.} & = n_1[n_1-n_2+ n_4-2n_5+n_6+u_3-u_2-4]+3n_1\\ &T_2^{grav.} & = n_2[n_1+n_3- n_4-n_5-n_6-u_1+u_4] \end{eqnarray} If non-Abelian anomaly cancellation equations (\ref{z12asim}) are used then terms inside brackets vanish and the correct contributions to be canceled by GS terms are obtained. The other cases proceed in the same way. \section{ Conclusions and remarks} We have studied the relationship between cancellation of tadpoles and anomalies in compact Type IIB $D=4$ vacua. We have found that only a subset of the tadpole conditions are in general needed in order to get anomaly cancellations. It is worth discussing what characterizes the twisted tadpoles whose cancellation is not required in order to get anomaly cancellation. In order to do that it is useful to do a T-duality transformation along the three complex compact dimensions. Now we have 3-branes and 7-branes (with their world-volume including the first two complex compact planes) instead of 9-branes and 5-branes. We can now decompactify and consider the $D=4$, $N=1$ field theory living in the intersection of the 3-branes and 7-branes. Now, in this non-compact orientifold one has to impose the tadpole conditions corresponding to a given twist {\it only if} the flux of the RR charge originating at the fixed point cannot escape to infinity. For example, as we discussed in section 3, in the $Z_3\times Z_3$ orientifold one has to impose the tadpole condition only for the twist $v=\frac 13(1,1,-2)$, because for twists leaving one fixed complex plane like e.g., $v=\frac 13(1,-1,0)$ the RR flux can escape to infinity. In the $Z_{12}$ case, which is another case in which a particular tadpole equation (\ref{tcup}) is not needed to get anomaly cancellation, something similar happens. In this case the twist is of the form $v=\frac 14(1,-1,0)$ and the RR flux can escape to infinity (in the non-compact case) through the third complex plane which is transverse to the D7-brane worldvolume. Now, coming back to the {\it compact} orientifold case, since the massless charged spectrum from open strings is the same for both compact and non-compact cases, the models are still going to be anomaly-free. Thus the extra tadpole equations needed in the compact case will not be related to anomaly cancellation. The $Z_6'$ orientifold example shows again how things go. In this case one would be tempted to say that tadpoles associated to the twist $2v=1/3(1,0,-1)$ should not be needed for anomaly cancellation, since the RR-flux could escape through the second complex plane. But that is not the case because the plane which is transverse to the D7-branes is the third one, which is rotated, no flux can escape through it. From the $D=4$ effective field theory point of view a natural question emerges. If there are tadpole constraints which are not needed for gauge anomaly cancellation, what is the low-energy symmetry which is guaranteed by them? We do not have a definite answer to that question but it seems sensible to believe that some other type of symmetries are guaranteed by them. In particular all these models have sigma-model $U(1)$ symmetries as well as discrete gauge symmetries and cancellation of their anomalies could require the extra constraints \cite{iru2}. Let us finally comment on the $D=6$ case. In $D=6$, $N=1$ models, anomaly and tadpole cancellation conditions are totally equivalent since there are no twists with unrotated complex planes and no possibility for the RR flux to escape. Indeed, an explicit analysis along the lines we described for the $D=4$ case gives rise to this equivalence. Concerning new solutions to the tadpole cancellation conditions, it is easy to see that the $Z_2$, $Z_3$ and $Z_4$ $D=6$, $N=1$ models of \cite{gj,dp2} have as unique solutions the ones given in those references. However the $Z_6$ orientifold admits more solutions than those reported there. The reason for this, as it happened in the $D=4$ case is that the $Z_6$ twist (unlike the other three cases) has only one fixed point, the one at the origin. Thus the RR flux due to D9-branes can partially cancel that from D5-branes and lead to new solutions. The general group one can obtain is the same as that in (\ref{ggz6}). \bigskip \bigskip \bigskip \centerline{\bf Acknowledgements} We are grateful to A. Font, R.~Rabad\'an and G. Violero for useful discussions. G.A. thanks Departamento de F\'{\i}sica Teorica at UAM for hospitality and financial support. A.~M.~U. is grateful to M.~Gonz\'alez for encouragement and support. L.E.I. thank CICYT (Spain) and the European Commission (grant ERBFMRX-CT96-0045) for financial support. The work of A.~M.~U. is supported by the Ram\'on Areces Foundation (Spain). \newpage
\section{Introduction} It is now 45 years since Cocconi \cite{kn:cocconi} drew attention to the broadening effects of the geomagnetic field on the lateral development of electron-photon cascades in the atmosphere. Allen \cite{kn:allen} discussed the interaction of the geomagnetic field and the cascade electrons/positrons as the origin of the radio frequency emission produced by large cosmic ray showers. Earnshaw {\it et al} \cite{kn:earnshaw} suggested that the geomagnetic separation of muons could be used to estimate the height of origin of showers. Porter \cite{kn:porter} and Browning and Turver \cite{kn:brover} made calculations for the production of \v{C}erenkov radiation by electron-photon cascades in the atmosphere, with allowance for the effects of the geomagnetic field. Ground based gamma ray astronomy exploits the magnifying effects of the atmosphere to enable the detection of very low fluxes of gamma rays from a variety of sources \cite{kn:fegan97}. The cascade of electrons initiated by a primary gamma ray produces, at an altitude of about 10 km, \v{C}erenkov light which reaches the ground in a pool $\sim 300$ m in diameter. It is possible to detect gamma rays of energy $>$ 300 GeV via this \v{C}erenkov light produced in the atmosphere with an effective collection area, for our Mark 6 telescope, of $5 \times10^{4}$ m$^{2}$. Bowden {\it et al} \cite{kn:bowden91} discussed the effect of the geomagnetic field on the performance of ground based gamma ray telescopes. The interaction of the field and the cascade electrons produces a broadening of the atmospheric \v{C}erenkov light image resulting in a reduction in the density of light sampled by the telescope; so the energy threshold for the telescope increases. This is observed in the higher count rate for a telescope detecting showers propagating along the lines of the field (with no spreading) than when observing cascades developing perpendicular to the field lines (and being spread), all other factors being the same. Typical differences in measured count rate were about 20\% in these extreme cases. The possibility was noted, on the basis of simulations, that an associated rotation of the direction of the \v{C}erenkov light images may occur for cascades developing under high magnetic fields and in unfavourable directions. This could be of importance in ground-based gamma ray studies since the orientation of the image in gamma ray initiated cascades is the key to rejection of $>$ 99\% of the charged cosmic ray background \cite{kn:hillas85}. Lang {\it et al} \cite{kn:lang93} showed that the measurements of TeV gamma rays from the Crab nebula using the imaging \v{C}erenkov technique were not significantly affected when the magnetic field was $<$ 0.35 G. We report measurements made using a ground based gamma ray telescope of cascades developing under the influence of fields up to 0.55 G. Observations with the Mark 6 telescope operating in Narrabri, Australia which is discussed by Armstrong {\it et al} \cite{kn:armstrong} are subject to such magnetic fields when observing objects to the south. The observational data demonstrate all the effects of the geomagnetic field on cascades which are predicted by simulations. \section{Geomagnetic effects on EM showers} \subsection{Simple model} The lateral distribution of electrons and positrons near the maximum of an electromagnetic shower in air is approximately axially symmetric. The presence of a transverse component of magnetic field will cause the electrons and positrons to separate and the lateral distribution to become wider along the radial direction perpendicular to the transverse field component. The order of magnitude of this effect may be estimated. A relativistic electron of energy $E$ (MeV) will follow a circular path in a transverse magnetic field of strength $H$ (G) with a radius of curvature $R$ (km) of $\sim{E}/(30 H)$. The energy of the typical \v{C}erenkov light-producing electron at shower maximum is $\sim$ 100 MeV (the critical energy in air), which is at a height of $\sim$ 10 km for a 250 GeV primary gamma ray. The radius of curvature of these electrons for a maximum transverse magnetic field of 0.56 G is $\sim$ 6 km. The typical length of such an electron's path in air at 10 km altitude is $\sim$ 1 km, giving a lateral deflection of $\sim$ 80 m. At a height of 10 km this corresponds to an angular deflection of $\sim 0.5 ^{\circ}$, the same order as the Coulomb scattering width. It should be possible to detect this as a broadening of the shower along one axis. The effect of the broadening will depend on the orientation of the image with respect to the direction of the magnetic field. The direction of the shower axis in the atmosphere is within $\sim$1$^{\circ}$ of the telescope-source line. The transverse magnetic field component therefore lies very close to the plane of the focal image of the shower. The parameters of an image which are used to discriminate between gamma rays and hadrons depend primarily on the second spatial moments. In a coordinate frame in which the x-axis is aligned with the projected field direction, these are $\sigma _{\rm x}^{2}$ , $\,\sigma _{\rm y}^{2}$ and $\sigma _{\rm xy}$. The orientation angle of the long axis of the image is \mbox{$\tan ^{-1} \left[\left(\sigma _{\rm y}^{2}-\sigma _{\rm x}^{2} + \sqrt{(\sigma _{\rm y}^{2}-\sigma _{\rm x}^{2})^{2}+4\sigma _{\rm xy}^{2}}\right)/(2\sigma_{\rm xy})\right]$}. Those shower images whose major axes are aligned with the projected direction of the magnetic field have \mbox{$\sigma _{\rm xy}=0$}. The moment $\sigma _{\rm y}^{2}$ will be expected to be increased by the magnetic field and the images will be widened only. For images at other orientations, \mbox{$\sigma _{\rm xy} \neq 0$} and the result of an increase in $|\sigma _{\rm xy}| $ will be to rotate their minor axes towards the projected field direction. The main parameter used to discriminate gamma rays is {\it ALPHA}, the angle contained within the long axis of the image and the radius vector from the source position to the centroid of the image. Any rotation of the image may affect {\it ALPHA} and hence the sensitivity for gamma ray detection. This is discussed further in section~\ref{simulations}. \subsection{Simulations}\label{simulations} Monte Carlo simulations of gamma ray and hadron cascades have been made to investigate the strength of this geomagnetic effect. These simulations have been performed using a Monte Carlo code developed from that used in our earlier work \cite{kn:bowden91} which incorporates the responses of the University of Durham telescopes. The results of the simulations have been validated against established simulation codes e.g. MOCCA \cite{kn:hillas85}. Agreement has been found in all cases. The consequence of an increase in the lateral and angular spread of the shower electrons is a decrease in the density of \v{C}erenkov light on the ground. This leads to a decrease in counting rate for fixed threshold, which was reported by us from measurements, using a non-imaging telescope, of counting rate against azimuth at fixed zenith angle \cite{kn:bowden91}. Figure~\ref{fig:simtotal} shows the simulated reduction in the counting rate of a gamma ray telescope of the type used by us for hadron primaries, relative to the zero-field case, caused by a transverse magnetic field of 0.5 G. This is consistent with our earlier measurements --- a reduction of up to $\sim 20\%$ in count rate for gamma ray energies $\geq 300$ GeV \cite{kn:bowden91} and, apparently, a greater reduction for $< 300$ GeV cascades. \begin{figure}[tb] \centerline{\psfig{file=chadwick01.EPS,height=6cm}} \caption{The simulated effect of the geomagnetic field on the brightness of the observed image. We show the \% reduction in event rate for hadrons as a function of primary energy for a field of 0.5 G relative to the zero-field case.} \label{fig:simtotal} \end{figure} In our earlier work \cite{kn:bowden91} we suggested that the distribution of recorded photons in an imaging camera may be distorted. We show in figure \ref{fig:simimage} the pattern of photons in individual simulated 1 TeV gamma ray cascades developing under the effects of 0 and 0.56 G geomagnetic fields. We also show the predicted responses of our camera after allowing for optical distortion, pixellation of the camera, measurement noise and passing through our standard data analysis package. The simulations for the pairs of cascades developing under the influence of a field and no field have been initiated using the same random number seed; this ensures, as far as is possible, that the early development of the cascade (which dominates the resulting image and is nearly independent of field) is identical for the two simulations. The expected effects on the shape of the image (widening or lengthening and rotation, depending on the orientation of the image with respect to the magnetic field direction) was apparent, both in the distribution of photons hitting the detector and in the processed image. \begin{figure}[tb] \centerline{\psfig{file=chadwick02.eps,height=15cm}} \caption{Simulations of a 1 TeV gamma ray cascade. The simulation has been performed with geomagnetic field values of (a) 0 and (b) 0.56 G. For each image, both a raw photon map (upper plot) and the results of fitting standard image parameters after allowance for mirror defects, pixelation and noise (lower plot) are shown.} \label{fig:simimage} \end{figure} The predicted width of a shower image was investigated as a function of magnetic field strength and angle for gamma rays of energy 500 GeV. Results are shown in figure~\ref{fig:simswid} for transverse field strengths of 0.0, 0.2 and 0.5 G. No allowance is made for the broadening effects of the defects in the mirror or the pixellation of the recording PMT camera. The effect of a 0.5 G field on the width (minor axis) of such undistorted gamma ray images is seen to be greatest when the image minor axis is perpendicular to the projected field direction, as expected. \begin{figure}[bt] \centerline{\psfig{file=chadwick03.eps,height=6cm}} \caption{The effect of the geomagnetic field on the width of simulated 500 GeV gamma ray shower images which are not subjected to any broadening effects of the mirror or camera. $\fullcircle$ is for a geomagnetic field of 0.0 G, $\opencircle$ for 0.2 G and $\blacktriangledown$ for 0.5 G.} \label{fig:simswid} \end{figure} Small transverse fields of 0.2 G have a limited effect on the image width. However, the effect on the pointing angle {\it ALPHA} become noticable and may vary both with the image's angle to the magnetic field and with its eccentricity (defined as the ratio {\em WIDTH}/{\em LENGTH}). For example, an image with very low eccentricity may be widened only by a small amount but rotated through a large angle. Histograms in {\it ALPHA} are shown in figure~\ref{fig:simsalpha} for the simulated gamma rays, for which widths were shown in figure~\ref{fig:simswid}, under the influence of fields of 0.0, 0.2 and 0.5 G. \begin{figure}[tb] \centerline{\psfig{file=chadwick04.eps,height=8cm}} \caption{The effect of the geomagnetic field on the distribution in pointing angle {\it ALPHA} for simulated 500 GeV gamma ray shower images. (a) is for a geomagnetic field of 0.0 G, (b) for 0.2 G and (c) for 0.5 G. Again, there is no allowance for the effects of mirror optics or camera pixellation.} \label{fig:simsalpha} \end{figure} The effects of mirror quality and camera pixellation of a typical telescope may reduce this change in width but may not mask the change in orientation of the image. \subsection{Magnetic fields appropriate to Narrabri observations} \begin{figure}[tb] \centerline{\psfig{file=chadwick05.ps,height=6cm}} \caption{The tracks for various potential gamma ray sources plotted to show the geomagnetic fields experienced at Narrabri, Australia.} \label{fig:tracks} \end{figure} We show in figure~\ref{fig:tracks} the celestial sphere, as viewed from the site of the Mark 6 telescope in Narrabri, Australia, showing the loci of constant magnetic field strength. We indicate the tracks on the sky for the telescope during observations of a number of potential gamma ray sources which have been intensively studied. In fact, most of the sources studied at Narrabri are in the azimuth range $135^\circ - 225^\circ$ and in directions for which the magnetic field is $>$ 0.35 G and in many cases $>$ 0.5 G. The telescope sensitivity is therefore likely to be reduced by the broadening of the distribution of {\it ALPHA} for many of our observations, unless corrections are applied. \section{The equipment} All data reported here have been obtained with the Mark 6 VHE gamma ray telescope which has been described in detail by Armstrong {\it et al} \cite{kn:armstrong}. The telescope measures gamma rays in the energy range 300 GeV -- 10 TeV, with a 50\% trigger probability at an energy of $\sim 300$ GeV. It comprises an alt-azimuth mount with three 7m diameter f/1.0 aluminium mirrors with an rms spread for a point source of $0.18^\circ$. The pixel size for a $1"$ diameter photomultiplier is $0.25^\circ$. The focal plane of the central mirror contains a 109-pixel camera comprising a close-packed hexagonal array of 91 1-inch diameter photomultipliers surrounded by a ring of 18 2-inch diameter photomultipliers. The field of view of the imaging camera is $3^\circ$ wide. The mirrors on the left and right of the mount each contain a triggering detector of 19 2-inch photomultipliers in a close-packed hexagonal array. Each photomultiplier is contained in a mumetal shell designed to give magnetic and electrostatic shielding. The effect of the geomagnetic field on the performance of the photomultipliers, with and without shielding, has been studied \cite{kn:roberts}. For the 1-inch Hamamatsu R1924 PMT the effect of a 0.5 G field on the gain was not measureable and less than 1\%. For the 2-inch Phillips XP3422 PMT the measured decrease in gain due to a 0.5 G field was 1\%. Shielded photomultipliers of both types show no measurable effects due to the magnetic field. The control of the attitude of the telescope is via DC servomotors driving onto gears mounted directly on the telescope structure. Angles are sensed by absolute digital 14-bit shaft encoders with a resolution of $0.022^{\circ}$. The calculated azimuth and zenith of a source is compared by a digital servomechanism (employing 12-bit resolution) with the shaft encoder outputs at 100 ms intervals. The error signals are passed via DACs to the DC motor amplifiers. These provide damping on acceleration and stabilise the movements of the telescope structure. Thus although the telescope pointing is known to a resolution of $0.022^\circ$, the source can be offset from the camera centre by up to $0.1^\circ$ by this mechanism. The attitude of the telescope is measured in two ways. The shaft encoder positions are recorded for each event to 14-bit accuracy. This gives a measurement of the pointing to $\pm 0.022^{\circ}$ within the telescope system. In addition, a coaxial optical CCD camera is mounted on the telescope. The output of this CCD camera is continuously monitored by microcomputer, which measures the position and brightness of a nominated guide star within the $ 2^\circ \times 2^\circ$ field. This information is integrated into the data stream on an event-by-event basis. Guide stars of magnitude $m_{\rm{v}} \leq 6$ can be employed, providing absolute position sensing for the telescope to better than $0.008^{\circ}$. False source analysis has been shown to be a useful method of demonstrating that $\gamma$-ray like events originate from the source direction \cite{kn:kifune1995}. Conversely, such analyses may also be used to verify the steering performance of a telescope, if an established $\gamma$-ray emitter is observed. However, we have only one data set for a gamma ray source containing data which are {\em not} subject to the effects of a strong geomagnetic field --- see section \ref{pks2155alpha}. \section{Experimental results} \subsection{Data} Most of the data considered here were taken during routine observations of potential gamma ray sources during 1996 -- 1998. A small amount of data was taken in dedicated measurements at fixed values of azimuth and zenith, corresponding to a range of values of the geomagnetic field. The only restriction imposed on data was that events should lie within $1^\circ$ of the centre of the camera (to avoid edge effects) and were large enough (5 times the triggering threshold) to ensure that their shape was well measured. \subsection{ Reduction in image brightness} The investigation of the sensitivity of the count rate to the magnetic field which established the effect with the Mark 3 and Mark 4 telescopes \cite{kn:bowden91} has been repeated for the Mark 6 telescope. We find that the count rate observed with the Mark 6 telescope for cascades developing under minimum and maximum values of the geomagnetic field at a constant zenith angle of $40^\circ$ differs by about 15\%, as expected. \subsection{Effects on the shape of \v{C}erenkov images of hadron showers} The widths of images recorded by the Mark 6 telescope in directions parallel and perpendicular to the magnetic field have been investigated. For a large sample of cascades measured during observations of a range of sources the differences of the mean width in the parallel and perpendicular directions are shown in figure \ref{fig:distortion} for a range of values of distorting field. The widths of images in directions perpendicular to the field are significantly greater than those in parallel directions for fields $> 0.4$ G. Although the differences in widths are small, because of the effects of pixellation and noise which are common to all data, the measurements are free from systematic effects. This is because the orientation of images to the magnetic field depends on the orientation of the image in the camera. Measurements of the width of the images parallel and perpendicular to the magnetic field are derived from events distributed throughout the same observation. We have attempted to measure the differences in widths of images in cascades measured at fixed zenith angles of $40^\circ$ for values of fields less than and greater than 0.35 G. Such measurements are difficult and require large exposures ($\gtrsim 100$ hr) to give significant differences. For example, with exposures $\sim 1$ hr we find for $B < 0.35$ G, the mean width is $0.292^\circ \pm 0.0035^\circ$ (SD $= 0.041^\circ$) and for $B > 0.35$ G the mean width is $0.283^\circ \pm 0.0036^\circ$ (SD $= 0.042^\circ$). The small difference in means, significant at the $\sim 2 \sigma$ level, demonstrates the difficulty in making such a small measurement for which the standard deviation is approximately similar for all $B$. \begin{figure}[tb] \centerline{\psfig{file=chadwick06.eps,height=6cm}} \caption{The observed difference in the mean width ($D_{\rm width}$) for images in hadronic cacades developing parallel and perpendicular to fields of varying strength for zenith angles in the range $35^\circ - 40^\circ$. The concentration of data at values of high magnetic field strength reflects the directions of observation of potential VHE sources from Narrabri.} \label{fig:distortion} \end{figure} \subsection{Effects on the orientation of \v{C}erenkov images of hadron showers} In the absence of any magnetic field and detector triggering biases, the distribution of the orientation of the images in the camera of a telescope will be isotropic. If we define a coordinate system where the angle $a_{\rm x}$ is the angle of the long axis of an image to the horizontal in the camera frame, then the distribution of $a_{\rm x}$ should be flat between $-90^{\circ}$ and $90 ^{\circ}$. Minor deviations from uniformity in angle might be expected near to threshold because of the increased effect of small changes in triggering probability and the six-fold symmetry of the hexagonal arrangement of close-packed detectors. \subsubsection{Distribution of projected shower angles} We show in figure~\ref{fig:aysmallh}(a) the distribution in the angle $a_{\rm x}$ for background hadronic events which are subject to small values of the transverse magnetic field (0.15 G). We show in figure~\ref{fig:aysmallh}(b) the distribution in $a_{\rm x}$ for images due to hadron-induced cascades recorded in special observations at fixed azimuth and zenith angles which gave a transverse geomagnetic field of 0.52 G. All of these events were recorded in dedicated observations with the telescope at a fixed zenith angle ($40^\circ$). Data at different values of the transverse magnetic field were obtained by varying the azimuth angle. All data presented here were recorded within a period of one hour, so potential variations due to changes in camera performance, atmospheric clarity, etc. were minimised. The data were processed following our normal procedures (see e.g. \cite{kn:chadwick98b}). The requirement that images have a minimum brightness and fall within the camera ensures that the data are free of effects of variations in night sky brightness. The data for $B = 0.15$ G show the expected distribution indicative of a near isotropic distribution of directions in the camera with a small peak. We note a strong anisotropy, with a peak containing twice the number of events resulting from the skewing or distortion of the image, for events subject to a $0.5$ G field. The maximum of the distortion occurs at $a_{\rm x} = 0$, which is appropriate for those observations which were made at magnetic south. \begin{figure}[tb] \centerline{\psfig{file=chadwick07.eps,height=6cm}} \caption{Distribution of $a_{\rm x}$ for cascades developing under (a) a small transverse geomagnetic field ($B = 0.15$ G) and (b) a large transverse geomagnetic field ($B = 0.52$ G). \label{fig:aysmallh}} \end{figure} \subsubsection{Correlation of amplitude of $a_{\rm x}$ peak and geomagnetic field strength} The magnitude of the anisotropy displayed in figure~\ref{fig:aysmallh}(b) should depend on the strength of the magnetic field. In figure~\ref{fig:ayampvh} we plot the amplitude of the peak of the distortion in the $a_{\rm x}$ distribution as a function of the magnetic field strength. Each point corresponds to the result for a 15 minute segment of data. \begin{figure}[tb] \centerline{\psfig{file=chadwick08.EPS,height=6cm}} \caption{The measured amplitude of $a_{\rm x}$ anisotropy as a function of transverse geomagnetic field.} \label{fig:ayampvh} \end{figure} Note that for values of transverse component of the geomagnetic field less than 0.35 G there is no great distortion of the $a_{\rm x}$ distribution, as suggested by the work of Lang {\it et al} \cite{kn:lang93}, but for values of field in excess of 0.4 G substantial distortion occurs. \subsubsection{Correlation of directions of $a_{\rm x}$ peak and geomagnetic field} The position of the peak in the $a_{\rm x}$ distribution depends on the angle between the projected magnetic field and the vertical direction in the camera --- $H_{\rm FOV}$. We show in figure~\ref{fig:aydirvh} the correlation between $H_{\rm FOV}$ and the position of the peak in $a_{\rm x}$. The data demonstrate the expected relation between these angles. Again, each point is based on a 15 min sample of data taken during routine observations of a range of potential gamma ray sources. \begin{figure}[tb] \centerline{\psfig{file=chadwick09.eps,height=6cm}} \caption{The measured direction of $a_{\rm x}$ anisotropy as a function of the direction of transverse geomagnetic field. $\opentriangledown$ represents data from the directions of Cen X-3, $\blacktriangledown$ from SMC X-1, $\opencircle$ from PKS 2005--489 and $\fullcircle$ from PKS 2155--304 to cover a range of $H_{\rm FOV}$.} \label{fig:aydirvh} \end{figure} \subsection{Summary} We have evidence that the geomagnetic field should, and does, influence the lateral development of atmospheric cascades. The observed effects on the threshold of a telescope and on the shape of the image and the magnitude and phase of the distortion of the pointing of images produced by background cosmic ray protons are as expected. \section{The orientation of gamma ray images} A number of potential gamma ray sources has been observed using the Mark 6 telescope at Narrabri. In several cases there is evidence for gamma ray emission \cite{kn:chadwick98a,kn:chadwick98b,kn:chadwick99}. The evidence comes mainly from a comparison of the {\it ALPHA} distributions for data selected on the basis of image shape for the ON-source and OFF-source scans. The difference between the {\it ALPHA} distributions should show an excess of events --- the gamma ray candidates --- at low values of {\it ALPHA}. If the data were taken in geomagnetically unfavourable directions, as is the case for most of our data, it might be expected that the {\it ALPHA} distribution of the excess events would be wider than that for data taken in more favourable geomagnetic directions. Consideration of data recorded from a gamma ray source which are subject to $B < 0.35$ G may provide the only true indication of the {\it ALPHA}-distribution for our telescope. \subsection{Observations of PKS 2155--304}\label{pks2155alpha} Observations of this object were made with the cascades recorded over a range of transverse geomagnetic field strengths between 0.25 and 0.5 G. The total data set contained 41 hrs of observation, as reported \cite{kn:chadwick99}. The {\it ALPHA} plot for the difference between the ON-source and OFF-source data taken at zenith angles $\theta < 45^{\circ}$ is shown in figure~\ref{fig:alphaplot2}(a). The significance of the excess at {\it ALPHA} $< 30^\circ$ is $5.7 \sigma$ according to the maximum likelihood method \cite{kn:gibson82}. \begin{figure}[tb] \centerline{\psfig{file=chadwick10.eps,height=6cm}} \caption{The distribution in {\it ALPHA} of excess gamma ray events from PKS 2155--304. (a) is for all events, (b) those for transverse fields $<$ 0.35 G. and (c) those for transverse fields $>$ 0.35 G.} \label{fig:alphaplot2} \end{figure} We are able to select a subset of data for which the strength of the projected field to which the cascades were exposed was $<$ 0.35 G and for which minimal distorting effect would be expected. The {\it ALPHA} plot for this subset is shown in figure~\ref{fig:alphaplot2}(b). The excess events all have {\it ALPHA} $ < 15^\circ$ (significance $4.7 \sigma$). The distribution is narrower than that of the total dataset and is typical of that expected for a $0.25 ^{\circ}$ pixel camera. (In the absence of any other data for gamma rays detected with our telescope and {\it not} subject to the effects of the magnetic field, we assume that this is reasonable.) The {\it ALPHA} plot for the majority of the events for which the field is $>$ 0.35 G, is shown in figure~\ref{fig:alphaplot2}(c). It is evident that the width of the peak is larger for these events recorded under the influence of higher transverse magnetic fields, with equal populations for values of {\it ALPHA} between $0^\circ$ -- $15^\circ$ ($3.6 \sigma$) and $15^\circ$ -- $30^\circ$ ($3.1 \sigma$). It should be noted that the peak at $60^\circ$ -- $70^\circ$ is superimposed on an {\em ALPHA}-plot with increasing frequency for {\em ALPHA} approaching $90^\circ$ --- see \cite{kn:chadwick99}. The significance of this peak is therefore $\sim 2 \sigma$ (before allowing for the number of bins in the {\em ALPHA}-plots). On the basis of our only measurement of a gamma ray source not subject to the effects of the geomagnetic field, we have evidence that the excess events with $15^\circ <$ {\em ALPHA} $< 30^\circ$ are confined to those measurements made with fields $> 0.35$ G. This conclusion is significant at the 2.5\% level, on the basis of a $2 \times 2$ contingency test of the populations of the $15^\circ - 30^\circ$ bins in figures~\ref{fig:alphaplot2}(b) and \ref{fig:alphaplot2}(c). \section{Conclusions} We have demonstrated that, on the basis of simulations and measurement, the \v{C}erenkov images from gamma rays and cosmic rays are broadened and rotated by the geomagnetic field. The broadening results in an increase in the telescope threshold and a reduction in counting rate. The rotation of the images away from the projected direction of the magnetic field in the image plane broadens the {\it ALPHA} distribution. These results suggest that the geomagnetic field can have an important effect on the operation of atmospheric \v{C}erenkov telescopes in some directions and that, for detected and candidate sources, our Narrabri site is particularly susceptible. These effects may be removed using appropriate correction techniques; unlike noise, geomagnetic effects do not reduce the information contained in the image. \ack We are grateful to the UK Particle Physics and Astronomy Research Council for support of the project and the University of Sydney for the lease of the Narrabri site. The Mark 6 telescope was designed and constructed with the assistance of the staff of the Physics Department, University of Durham. \section*{References}
\section{Introduction} Systematic investigations of A($e,e'p$) reactions at Saclay, NIKHEF, Mainz and Bates have produced an impressive amount of data for various target nuclei \cite{kelly96}. For the sake of minimizing the uncertainties with respect to the reaction mechanism, a large fraction of these investigations were done under or close to quasi-elastic kinematics (Bjorken $x= \frac {-q ^\mu q _ \mu } {2M_N \omega } \approx$~1). These data sets highlight at the same time the success and the limits of the independent-particle-model (IPM) for atomic nuclei \cite{vijay98}. Indeed, after corrections for final-state interactions and Coulomb distortions of the electron probe, the shape of the deduced proton momentum distributions are systematically in line with the predictions of modern formulations of the nuclear IPM. On the other hand, below the Fermi momentum the absolute magnitude of the deduced momentum distributions are systematically, i.e. independent of the nucleon momentum, lower than IPM predictions. To cut a long story short, the major conclusion from this world-wide $(e,e'p)$ effort seems that an appropriate non-relativistic picture of the nucleus is roughly compatible with {\sl 70\% mean-field behaviour and 30\% ``correlations''} an observation which is still frequently ignored in various nuclear structure calculations and model developments. The energy of the available electron beams with a large duty factor ($\epsilon \leq $~1~GeV) made that most of this aforementioned $(e,e'p)$ work was done at four-momentum transfers of the order $Q^2= -q ^\mu q _ \mu \leq$~0.2~(GeV/c)$^2$. With the advent of the TJNAF and an upgraded Mainz electron facility higher values of $Q^2$ come into reach of experimental exploration. Amongst the major physics' goals motivating exclusive ($e,e'p$) measurements from finite nuclei at higher momentum transfer ($Q^2 \geq 0.2~$ (GeV/c)$^2$) one can mention the following ones. The higher $Q^2$ conditions and unmistakingly smaller distance scales probed should make it feasible to achieve a better understanding of the short-range mechanisms in nuclei. At the same time, one could hope to find experimental evidence for the onset of quark and gluon degrees of freedom. Presumably, the most convincing evidence pointing into that direction could come from measurements for processes that are fairly well understood at lower $Q^2$ (like ($e,e'p$) at x=1) and turn out to be completely at odds with meson/baryon models when higher $Q^2$ regimes are entered. Another challenge for ($e,e'p$) measurements at higher momentum transfers is the question whether available relativistic models can succeed in producing a better agreement with the data sets than the non-relativistic ones and if so, which dynamical degrees of freedom make them to be substantially different from what is commonly implemented in the non-relativistic nuclear many-body models. Another fundamental question that has received a great share of attention for many years, is the question whether the nucleon properties (like electromagnetic form factors e.g.) are modified in the nuclear medium. This information is of unvaluable importance for models that embark on the ambitious program of understanding nuclei in terms of quark and gluon degrees of freedom \cite{lu98a,lu98b}. A challenging but at the same time rather ambiguity-free way of probing the medium-dependent form factors, are double polarization observables from A($\vec{e},e'\vec{p}$) measurements \cite{laget94,e89033,e93049,e91006}. Indeed, double polarization observables are conceived to be rather insensitive to ambiguities with respect to the final state interactions (FSI) that affect most of the other A($e,e'p$) observables. Other effects that are recognized to possibly complicate the interpretation of double polarization observables in terms of medium-dependent nucleon form factors are gauge ambiguities, channel-couplings and two-body current effects. Whereas the first two sources of possible dilutions were extensively studied by Kelly \cite{kelly1,kelly2} and found to produce very small corrections, the two-body current effects are not well studied for finite nuclei. Earlier efforts to study the role of two-nucleon currents in exclusive A($e,e'p$) reactions from finite nuclei (A$\geq$4) include the pioneering work of Suzuki \cite{suzuki89}, the systematic investigations by the Pavia group \cite{boffi90,boffi90a,boffi92} and the $^4$He($e,e'p$) studies as e.g. reported in Refs.~\cite{epstein93,laget94}. In order to make their calculations computationally more attractive, the results of the Pavia group were obtained with a two-body current operator that was formally reduced to an effective one-body one. This approximation adopts a Fermi-gas picture for the residual nucleus that allows to integrate out the coordinates of the second nucleon that gets involved in the photoabsorption process. For the work presented here, we treat the two-body currents in their full (non-relativistic) complexity and deal with the multi-dimensional integrals that automatically occur when several nucleons get involved in the (virtual) photoabsorption process. A similar sort of exact treatment was earlier adopted in our work reported in References \cite{veerle1,veerle2}. There it was found that the effect of the two-body currents on the A($e,e'p$) cross sections, just as the role of the coupled-channel effects, is gradually decreasing with increasing momentum transfer. This conforms to the findings with respect to the momentum-transfer dependence of the two-body current effects in the d($e,e'p$)n and $^4$He($e,e'p$) \cite{laget94}. Nevertheless, even at the highest momentum transfer considered in Ref.~\cite{veerle1} (q=600~MeV/c) the cross sections were predicted to be substantially affected by the two-body currents. In this work the role of meson-exchange and isobar degrees of freedom in A($\vec{e},e'\vec{p}$) and A($e,e'n$) is investigated in a wide range of four-momentum transfers (0.2 $\le Q^2 \le$0.8~(GeV/c)$^2$). Special emphasis is placed on the recoil polarization observables as they open perspectives to investigate possible medium modifications of the nucleon properties. The calculational framework will be introduced in Section~\ref{sec:model}. In Section~\ref{sec:observ} the adopted conventions for the ($\vec{e},e'\vec{p}$) observables will be introduced. The model for the bound and scattering states is described in Section~\ref{sec:fsi}. In Section~\ref{sec:current} the model assumptions with respect to the one- and two-body current operators will be summarized. The results of the calculations are contained in Section~\ref{sec:results}. Our conclusions are summarized in Section~\ref{sec:conclusions}. \section{Theoretical framework} \label{sec:model} \subsection{A($\vec{\mathrm{e}}$,e$'\vec{\mathrm{p}}$) observables and kinematics} \label{sec:observ} We consider processes in which a longitudinally polarized electron impinges on a nucleus and induces the following reaction \begin{equation} A \; + \vec{e} (\epsilon) \longrightarrow (A-1)(E_{A-1},\vec{p}_{A-1}) \; + N(E_N,\vec{p}_N) \; + e (\epsilon ') \; , \end{equation} to occur. For such a process the cross section reads in the one photon exchange approximation \begin{eqnarray} & & {d^5 \sigma \over d \Omega _N d \epsilon ' d \Omega _{\epsilon '}} (\overrightarrow{e},e'N) = {1 \over 4 (2\pi)^5 } p_N E_N f_{rec}^{-1} \sigma_{M} \nonumber \\ \times & & \Biggl[ v_T {W_T} + v_L {W_L} + v_{LT} {W_{LT}} + v_{TT} {W_{TT}} + h \biggl[ v'_{LT} {W'_{LT}} + v'_{TT} {W'_{TT}} \biggr] \Biggr] \; , \label{eq:eep} \end{eqnarray} where $f_{rec}$ is the recoil factor \begin{equation} f_{rec} = \left| 1 + \frac {E_N} {E_{A-1}} \left(1 - \frac {\vec{q} \cdot \vec{p}_N} {p_N^2} \right) \right| \; , \end{equation} and $\sigma _M$ the Mott cross section \begin{equation} \sigma _M = \frac {\alpha^2} {4 \epsilon ^2} \frac {cos ^2 \frac {\theta _e} {2}} {sin ^4 \frac {\theta _e} {2}} \: . \end{equation} The electron kinematics is contained in the kinematical factors \begin{eqnarray} v_T &=& tg^2 \frac{\theta_e}{2} - \frac{1}{2}\left(\frac{q_{\mu}q^{\mu}}{\vec{q}^2}\right) \\ v_L &=& \left( \frac{q_{\mu}}{\vec{q}} \right)^4 \\ v_{LT}&=&\frac{q_{\mu}q^{\mu}} {\sqrt{2} \mid \vec{q} \mid ^3} (\epsilon + \epsilon ') tg \frac{\theta_e}{2} \\ v_{TT}&=&\frac{q_{\mu}q^{\mu}}{2 \vec{q} ^2} \\ v'_{LT} & = & \frac {q_{\mu}q^{\mu}} {\sqrt{2} \vec{q}^2} tg \frac{\theta_e}{2} \\ v'_{TT} & = & \sqrt{\left( -\frac {q_{\mu}q^{\mu}} {\vec{q}^2} + tg ^2 \frac{\theta_e}{2} \right)} tg \frac {\theta_e} {2} \; , \end{eqnarray} whereas the structure functions are defined in the standard fashion \begin{equation} \begin{array}{ll} W_L = \left( J_0^{fi} \right)^* \left( J_0^{fi} \right) & W_T = \left( J_{+1}^{fi} \right)^* \left( J_{+1}^{fi} \right) + \left( J_{-1}^{fi} \right)^* \left( J_{-1}^{fi} \right) \\ W_{LT}= 2 {\cal R}e \left[ \left( J_{0}^{fi} \right)^* \left( J_{-1}^{fi} \right) - \left( J_{0}^{fi} \right)^* \left( J_{+1}^{fi} \right) \right] & W_{TT}= 2 {\cal R}e \left[ \left( J_{-1}^{fi} \right)^* \left ( J_{+1}^{fi} \right) \right] \\ W_{LT}^{'} = -2 {\cal R}e \left[ \left( J_{0}^{fi} \right)^* \left( J_{+1}^{fi} \right) + \left( J_{0}^{fi} \right)^* \left( J_{-1}^{fi} \right) \right] & W_{TT}^{'} = \left( J_{+1}^{fi} \right)^* \left( J_{+1}^{fi} \right) - \left( J_{-1}^{fi} \right)^* \left( J_{-1}^{fi} \right) \; . \\ \end{array} \end{equation} The above definitions for the structure functions and the kinematical variables correspond with those of Ref.~\cite{raskin}. We remind that apart from a negligible parity-violating component, the structure function W$'_{TT}$ vanishes identically and the W$'_{LT}$ in coplanar kinematics if no recoil polarization is determined. A new set of observables comes into reach of experimental exploration when performing polarimetry on the ejected hadron. This results in knowledge about the spin orientation of the ejectile and as e.g. illustrated in neutron form factor studies represents a powerful tool to address fundamental physical quantities. The formal framework for the electroproduction of polarized nucleons from nuclei is outlined in great detail in Refs.~\cite{raskin,pickle87,boffigiusti} and will not be repeated here. Here, we only review some basic concepts which mainly serves at introducing the conventions adopted. For the results presented below, the polarization of the escaping nucleon is expressed in the so-called barycentric reference frame that is defined by the following set of unit vectors (Fig.~\ref{fig:reffram}) \begin{eqnarray} \hat{\vec{l}} & = & \frac {\vec{p}_N} {\left| \vec{p}_N \right|} \\ \nonumber \hat{\vec{n}} & = & \frac {\vec{q} \times \vec{p}_N} {\left| \vec{q} \times \vec{p}_N \right|} \\ \nonumber \hat{\vec{t}} & = & \hat{\vec{n}} \times \hat{\vec{l}} \; . \end{eqnarray} Note that for coplanar kinematics $\hat{\vec{n}}$ determines the y axis of the reference frame. The escaping nucleon polarization observables can be determined through measuring {\bf ratios}. The {\sl induced polarization} can be addressed with unpolarized electrons (i=n,l,t) \begin{equation} {\mathrm P}_i = \frac {\sigma (s_{N}^{i}= \uparrow) - \sigma (s_{N}^{i}= \downarrow)} {\sigma (s_{N}^{i}= \uparrow) + \sigma (s_{N}^{i}= \downarrow)} \hspace{0.5cm} \; , \label{eq:indu} \end{equation} whereas the {\sl polarization transfer} also requires polarized electron beams (i=n,l,t) \begin{eqnarray} {\mathrm P}_i' & = & \frac { \left[ \sigma (h=1,s_{N}^{i}= \uparrow) - \sigma (h=-1,s_{N}^{i}= \uparrow) \right] - \left[ \sigma (h=1,s_{N}^{i}= \downarrow) - \sigma (h=-1,s_{N}^{i}= \downarrow) \right] } { \left[ \sigma (h=1,s_{N}^{i}= \uparrow) + \sigma (h=-1,s_{N}^{i}= \uparrow) \right] + \left[ \sigma (h=1,s_{N}^{i}= \downarrow) + \sigma (h=-1,s_{N}^{i}= \downarrow) \right] } \; , \label{eq:trans} \end{eqnarray} where {$s_{N}^{i}= \uparrow$} denotes that the ejected hadron is spin-polarized in the positive i direction (i=(n,l,t)) and $h$ is the helicity of the electron impinging on the target nucleus. $\sigma (h,s_{N}^{i})$ is a shorthand notation for the differential cross section for an electrodisintegration process initiated by an electron with helicity $h$ and for which the ejectile is detected with a spin polarization characterized by $s_{N}^{i}$. Throughout this work we adopt relativistic kinematics. \subsection{Final-state interactions} \label{sec:fsi} For the model calculations presented here, the bound and scattering states are produced by solving a Schr\"{o}dinger equation with a mean-field potential that is obtained from a Hartree-Fock calculation. The latter are performed with an effective nucleon-nucleon force of the Skyrme type. The mean-field potential includes central, spin-orbit and Coulomb terms. This model does not require any empirical input with respect to the inital and final-state potentials. Moreover, the orthogonality condition between the initial and final states is obeyed and gauge invariance is preserved at the one-body current level. A drawback of the model is that the inelastic processes, which are commonly accounted for through imaginary parts in the final-state potential, are only partially included. The major impact of the rescattering processes in exclusive ($e,e'p$) is generally conceived, however, to be a mere reduction of the absolute cross section. In the kinematics regime of interest here, the reduction factor can be estimated from nuclear transparency $(e,e'p)$ measurements as they were recently performed at TJNAF for various target nuclei \cite{potterfeld}. The differential cross section, the various structure functions and polarization observables are all calculated starting from the following set of transition matrix elements \begin{equation} \label{eq:matr} m^{fi} _ \lambda = <J_{R}M_{R} (E_x) ;\vec{p}_{N}, \frac {1} {2} s_{N}^{z} \mid J_{\lambda }(q)\mid J_{i}M_{i}> \; \; \; (\lambda = 0, \pm1) \; \end{equation} where $\mid \! J_{i}M_{i}\! >$ and $\mid \! J_{R}M_{R} (E_x)\! >$ refer to the quantum states of the target and residual A-1 nucleus and $s_{N}^{z}$ denotes the spin projection of the ejectile along the z-axis. The latter is chosen to coincide with the direction of the momentum transfer $\vec{q}$. When calculating the induced and transverse polarizations as defined in Eqs.~(\ref{eq:indu}) and ~(\ref{eq:trans}) the transition matrix element for the ejectile's spin pointing in a certain direction determined by the polar and azimuthal angle ($\theta ^*, \phi ^*$) are required. These can be easily obtained from the above matrix elements (\ref{eq:matr}) by remarking that \begin{eqnarray} <J_{R}M_{R} & & (E_x) \vec{p}_{N}, \frac {1} {2} s'_{N}(\theta ^* \phi ^*) \mid J_{\lambda }(q)\mid J_{i}M_{i}> = \\ \nonumber & & \sum _ {s_{N}^{z}} \left( {\cal D} _{s_{N}^{z} , s'_{N}} ^{(1/2)} \left( \theta ^*, \phi ^* \right) \right) ^* <J_{R}M_{R} (E_x) ;\vec{p}_{N}, \frac {1} {2} s_{N}^{z} \mid J_{\lambda }(q)\mid J_{i}M_{i}> \; \; \; (\lambda = 0, \pm1) \; , \end{eqnarray} where ${\cal D}^{(1/2)}$ is the Wigner ${\cal D}$-matrix for $j= \frac {1} {2}$ \begin{equation} {\cal D} ^{(1/2)} (\theta^*,\phi^*) = \left( \begin{array} {cc} cos \frac {\theta ^*}{2} & sin \frac {\theta ^*}{2} e^{i \phi ^{*}} \\ sin \frac {\theta ^*}{2} & -cos \frac {\theta ^*}{2} e^{i \phi ^{*}} \\ \end{array} \right) \; . \end{equation} In determining the matrix elements of Eq.~(\ref{eq:matr}), a standard multipole expansion for the electromagnetic current operators and the ``distorted'' outgoing nucleon wave is made. The multipole expansion of the current operators tends to converge more slowly as more extended systems and higher momentum transfers are addressed. For the highest momentum transfer considered here ($\mid \vec{q} \mid \approx 1$~GeV), convergence in the $^{16}$O calculations could only be reached after including all multipolarities up to $J_{max}$=30. This number is compatible with the estimate one would obtain by setting $J_{max}=2qR_A$, where $R_A$ is the nuclear radius determined by $R_A = 1.2 \; A^{1/3} \; \mathrm{(fm)}$. The large number of multipoles required for calculations at high momentum transfer is a serious numerical complication that prevents models that were developed for lower energy and momentum transfers from being easily extended to higher energy and momentum domains. Apart from the fact that a full treatment of the relativistic effects would be in order, the partial-wave expansion technique hampers extending the presented formalism beyond the range of momentum-transfers considered here ($\mid \vec{q} \mid \le$1~GeV/c). A quasi-elastic ($e,e'p$) process will predominantly excite those A-1 states $ \mid J_{R}M_{R} (E_x) >$ that bear a sizeable hole state component in their overlap with the ground-state of the target nucleus. This component of the final wave function can be constructed starting from the standard multipole expansion in terms of particle-hole $\left| p(lj\epsilon) h^{-1} \right>$ excitations out of the ground state of the target nucleus \cite{Ryc88,Mah} : \begin{eqnarray} \left| J_{R}M_{R} (E_x) ;\vec{p}_{N}, \frac {1} {2} s_{N}^{z} \right> & = & \sum_{lm_ljm} \sum_{JM} 4 \pi i^l \sqrt{\frac {\pi} {2 \mu p_N}} <j_h \; m_h \; j\; m \mid J\; M> \nonumber \\ & & \times <l \; m_l \; \frac{1}{2} \; s_{N}^{z} \mid j \; m> e^{i(\delta_l+\sigma_l)}Y_{lm_l}^{*}(\Omega_N) \mid p\left(lj\epsilon \right)h^{-1} \; ; \; JM > \;, \end{eqnarray} where $\epsilon \equiv p^2_N/(2 \mu)$, $\mu$ is the reduced mass of the outgoing nucleon, $\delta _l$ is the central phase shift and $\sigma _l$ is the Coulomb phase shift related to the electromagnetic part of the mean-field potential. The above expression has been derived for the following conventions with respect to the asymptotic behaviour of the continuum eigenstates $\mid p(lj\epsilon) > \equiv \varphi_{lj}(r,E)$ of the mean-field potential~ and the total A-body wave function: \begin{eqnarray} \varphi_{lj}(r,\epsilon) & \stackrel{r \gg R_A} {\longrightarrow} & \sqrt{\frac{2\mu} { \pi p_N}} \frac {sin(p_N r-\eta ln(2p_Nr) - \frac{\pi l}{2} + \delta _l + \sigma _l)} {r} \; , \nonumber \\ \left| J_{R}M_{R} (E_x) ;\vec{p}_{N}, \frac {1} {2} s_{N}^{z} \right> & \stackrel{r_A \gg R_A} {\longrightarrow} & \frac {1} {\sqrt{A}} \left( e^{i \vec{p}_N \cdot \vec{r}_A} + f_k (\theta) \frac {e^{ip_Nr_{A}}} {r_{A}} \right) \left| \frac{1} {2} s_{N}^{z} \right> (-1)^{j_h+m_h} c_{h-m_{h}} \mid J_i M_i > \; , \end{eqnarray} where $R_A$ is the nuclear radius. \subsection{One and Two-body current operators} \label{sec:current} The one-body current is derived from the on-shell covariant single-nucleon current \begin{equation} j^{\mu}= \overline u (\vec {p}_N,s_N^z) \Biggl[ F_1 (Q^2) \gamma ^{\mu} + F_2 (Q^2) \frac {i \sigma ^{\mu \nu} q _{\nu}} {2 M_N} \Biggr] u (\vec {p} _m,s_m^z) \; , \label{eq:dirac} \end{equation} where $F_1 (F_2)$ is the Dirac (Pauli) form factor. Even when making an abstraction from issues related to off-shell corrections, the derivation of an appropriate current operator with relativistic corrections to be used in calculations that adopt a non-relativistic description for the strong interaction part of the reaction process, is not free from ambiguities \cite{amaro98,fearing94,boffi98,jeschonn98}. For example, the two standard techniques to derive a non-relativistic reduction of relativistic hamiltonians, i.e. the Foldy-Wouthuysen (FW) and the ``direct Pauli reduction'' method, have been shown to make a difference as far as the final expressions for the current operators are concerned \cite{fearing94}. Admittedly, the problems encountered in implementing relativistic corrections in non-relativistic calculations can be avoided by adopting a fully relativistic description for the nuclear dynamics. Over the years, a number of fully relativistic models for ($e,e'p$) have been developed \cite{wright,udias,johans,mcdermott}. Basically, all of these models start from a relativistic formulation of the mean-field idea and do not embody current operators that go beyond the IA. Unfortunately, for finite nuclei a realistic, empirically well-founded and practicable relativistic nuclear-structure model that {\sl also accounts for the multi-nucleon degrees of freedom} is as yet not available. With this in mind we shall resort to a non-relativistic approach for the strong interaction dynamics of the $(e,e'p)$ process. We deem this model to be a reliable and valuable testing ground to estimate the relative importance of subnuclear degrees of freedom. The conventional way of deriving a nonrelativistic reduction of the above one-body current operator is the FW technique which relies on a ($q/M_N$) expansion. Recently, an alternative expansion in terms of $(p_m/M_N)$, with $p_m$ the momentum of the nucleon on which the absorption occurs, has been suggested \cite{amaro98,jeschonn98}. This expansion is based on the ``direct Pauli reduction'' method. In order $(q/M_N)^2$ and $(p_m/M_N)$ respectively, both methods give rise to a nonrelativistic four-current density operator that reads in coordinate space \begin{eqnarray} \rho(\vec{r}) & = & \sum _{i=1} ^{A} \Biggl( C_1(q,Q^2) G_E^i(Q^2) \delta (\vec{r} - \vec{r} _i) \nonumber \\ & & - C_2(q,Q^2) \frac {(2 G_M^i(Q^2) - G_E^i(Q^2))} {8 M_N^2 i} \vec{\nabla} \cdot \left[ \vec{\sigma} _i \times \left\{ \vec{\nabla}_i, \delta (\vec{r} - \vec{r}_i) \right\} \right] \Biggr) \nonumber \\ \vec{J}^{\perp}(\vec{r}) & = & \sum _{i=1} ^{A} C_3(q,Q^2) \Biggl( \frac {G_E^i(Q^2)} {2M_N i} \left\{ \vec{\nabla}_i, \delta (\vec{r} - \vec{r}_i) \right\} + \frac {G_M^i(Q^2)} {2M_N} \delta (\vec{r} - \vec{r}_i) \left( \vec{\sigma}_i \times \vec{\nabla} \right) \Biggr) \; , \label{eq:redcurrent} \end{eqnarray} where $\{ \dots, \dots \}$ denotes the anticommutator and it is implicitly understood that the operator $\vec{\nabla}$ acts only the photon field. The $G_E(Q^2)$ and $G_M(Q^2)$ are the Sachs form factors for which we have adopted the standard dipole form. The major difference between the current operators in the two earlier sketched approaches is not the operatorial form but the derived expression for the coefficients ($C_1,C_2,C_3$). Whereas with the FW method one finds \cite{roccojoe} \begin{equation} \begin{array}{ccc} C_1(q,Q^2) = \frac {1} {\sqrt{1+ \frac {Q^2} {4 M_N^2}}} \hspace{0.7cm} & C_2(q,Q^2) = 1 \hspace{0.7cm} & C_3(q,Q^2) = 1 \hspace{0.7cm} \; ,\\ \end{array} \end{equation} a $(p_m/M_N)$ expansion with the aid of the Pauli reduction technique leads to \cite{amaro98,jeschonn98} \begin{equation} \begin{array}{ccc} C_1(q,Q^2) = \frac {q} {\sqrt{Q^2}} \hspace{0.7cm} & C_2(q,Q^2) = \frac {1} {\sqrt{1+ \frac {Q^2} {4 M_N^2}}} \hspace{0.7cm} & C_3(q,Q^2) = \frac {\sqrt{Q^2}} {q} \hspace{0.7cm} \; .\\ \end{array} \end{equation} Unless otherwise specified all results of this paper are obtained with the four-current operator from Eq.(~\ref{eq:redcurrent}) using the $C$ coefficients as they are obtained with the FW method. The major difference in comparison with a strict non-relativistic approach is the introduction of a spin-orbit term $\rho ^{so} (\vec{r})$ in the charge density operator. This gives rise to a spin-orbit term in the Coulomb transition operator that determines the longitudinal strength \begin{equation} M_{JM}^{so}(q)= \int d \vec{r} j_J(qr) Y_{JM}(\Omega) \rho ^{so} (\vec{r}) \; . \end{equation} In coordinate space the one-body transition matrix element corresponding with this operator reads \begin{eqnarray} & & \left< n_a l_a \frac {1} {2} j_a \parallel M_{J}^{so}(q) \parallel n_b l_b \frac {1} {2} j_b \right> = \sum_{\eta \pm 1,J_2} (-1)^{\delta _{\eta,-1}+1} C_2(q,Q^2) \frac {(2 G_M(Q^2) - G_E(Q^2))} {8 M_N^2} \nonumber \\ & & \times 3 \sqrt{ 12(J+\delta_{\eta,+1})} (2J_2+1) \left\{ \begin{array}{ccc} J_2 & J + \eta & 1 \\ 1 & 1 & J \\ \end{array} \right\} \left\{ \begin{array}{ccc} l_a & \frac {1} {2} & j_a \\ l_b & \frac {1} {2} & j_b \\ J_2 & 1 & J \\ \end{array} \right\} \nonumber \\ & & \times \left< n_a l_a \parallel \left( \frac {d} {dr} j_J(qr) + (-1)^{\delta_{\eta,+1}} (J + \delta _{\eta,-1}) \frac {j_J(qr)} {r} \right) \left[ Y_{J+\eta} \otimes \left( \vec{\nabla} - \vec{\nabla}' \right) \right]^{(J_{2})} \parallel n_b l_b \right> \; . \end{eqnarray} In determining the two-body current operators we start from the observation that of all mesons carying the nucleon-nucleon interaction the pions play a predominant role. In our model calculations, the current operators that explicitly account for subnuclear $\pi$ and $\Delta$ degrees of freedom are effecively written in terms of the coordinates of the two nucleons involved. The meson-exchange and $\Delta$-isobar currents as they were implemented in our calculations have been discussed in detail in Ref.~\cite{jan97}. The two-body pion-exchange currents are derived from the one-pion exchange potential in the standard fashion. The derivation of the $\Delta$-current operator is somewhat more complicated. Indeed, this operator cannot be constrained through charge-current conservation and is therefore often referred to as a ``model-dependent'' operator \cite{roccojoe}. Moreover, of all pion-related two-body current contributions the $\Delta$-current is the only one that could be unambiguously shown to exhibit a quite strong medium dependence. We do not consider explicit $\Delta$ admixtures in the wave functions, so that all isobars in our model calculations are attached to a photon line. In our derivations, the $\pi N \Delta$ and $\gamma N \Delta$ coupling are considered in the standard form~: \begin{eqnarray} {\cal L} _{\pi N \Delta} & = & \frac {f_{\pi N \Delta}} {m _{\pi}} \left( \vec{S} ^{\dagger} . \vec{\nabla} \right) \left( \vec{T} ^{\dagger} . \vec{\pi} \right) \; , \\ {\cal L} _{\gamma N \Delta} & = & G_{\gamma N \Delta}(q_\mu^2) \frac {f_{\gamma N \Delta}} {m _{\pi}} \left( \vec{S} ^{\dagger} \times \vec{\nabla} \right) . \vec{A} \; \vec{T} _{z} ^{\dagger} \; , \end{eqnarray} where $\vec{S}$ and $\vec{T}$ denote the $\frac {1} {2} \rightarrow \frac {3} {2}$ spin and isospin transition operators, $\vec{A}$ is the external electromagnetic field and the coupling constants are $\frac {f^2_{\pi N \Delta}} {4 \pi}$ = 0.37 and $f_{\gamma N\Delta}$ = 0.12. The electromagnetic form factor of the delta $G_{\gamma N \Delta}(q_\mu^2)$ is parametrized as \cite{marcucci98} \begin{equation} G_{\gamma N \Delta}(q_\mu^2) = \frac {1} {\left( 1 - \frac {q_\mu^2} {\Lambda_1 ^2} \right)^2} \frac {1} {\sqrt{ 1 - \frac {q_\mu^2} {\Lambda_2 ^2} }} \; , \end{equation} where $\Lambda_1$=0.84~GeV/c and $\Lambda_2$=1.2~GeV/c. It is interesting to note that the $N-\Delta$ electromagnetic form factor $G_{\gamma N \Delta}$ is decreasing with $Q^2$ faster than the nucleon dipole form \cite{stoler93,frolov99}. With the above coupling lagrangians we arrive at the following expression for the $\Delta _{33}$-current in momentum space \begin{eqnarray} \label{eq:pidelta} &&\vec J_{\pi \Delta} (\vec q,\vec k_1,\vec k_2 ; \gamma N \Delta \longrightarrow NN) = {i \over 9}{{f_{\gamma N\Delta }f_{\pi NN}f_{\pi N\Delta }} \over {m_\pi ^3\,}} G_{\gamma N \Delta}(q_\mu^2) \nonumber\\ && \times \Biggl[ \left( G_{\Delta}^{I} + G_{\Delta}^{II} \right) \Biggl( 4 \vec \tau _{2,z} \left( \vec k_2 \times \vec q \right) {{\vec \sigma _2\cdot \vec k_2} \over {\vec{k_2}^2+m_\pi ^2}} + 4 {\vec \tau _{1,z}} \left( \vec k_1\times \vec q \right) {{\vec \sigma _1\cdot \vec k_1} \over {\vec{k_1}^2+m_\pi ^2}} \nonumber\\ && + (\vec \tau _1 \times \vec \tau _2)_z \left[ \left( {\vec \sigma _2 \times \vec k_1} \right) {{\vec \sigma _1\cdot \vec k_1} \over {\vec{k_1}^2+m_\pi ^2}} - \left( {\vec \sigma _1 \times \vec k_2} \right) {{\vec \sigma _2 \cdot \vec k_2} \over {\vec{k_2}^2+m_\pi ^2}} \right] \times \vec q \Biggr) \nonumber \\ && + \left( G_{\Delta}^{I} - G_{\Delta}^{II} \right) \Biggl( - 2i {\vec \tau _{2,z}} \left( \left( \vec \sigma _1 \times \vec k_2 \right) \times \vec q \right) {{\vec \sigma _2 \cdot \vec k_2} \over {\vec{k_2}^2+m_\pi ^2}} -2 i \vec \tau _{1,z} \left( \left( \vec \sigma _2\times \vec k_1 \right) \times \vec q \right) {{\vec \sigma _1\cdot \vec k_1} \over {\vec{k_1}^2+m_\pi ^2}} \nonumber\\ && -2i (\vec \tau _1\times \vec \tau _2)_z \left[ \vec k_2 {{\vec \sigma _2\cdot \vec k_2} \over {\vec{ k_2}^2+m_\pi ^2}} - \vec k_1 {{\vec \sigma _1\cdot \vec k_1} \over {\vec{ k_1}^2+m_\pi ^2}} \right] \times \vec q \Biggr) \Biggr] \;. \end{eqnarray} At the $\pi N \Delta$ and $\pi NN$ vertices, monopole form factors \begin{equation} \frac {\Lambda _ {\pi NN} ^2 - m _ \pi ^2} {\Lambda _ {\pi NN} ^2 + p _ \pi ^2} \end{equation} are introduced. They correct for finite size effects of the interacting baryons and regularize the $\pi NN$ interaction at extremely short distances. All results presented below are obtained with a cut-off parameter $\Lambda _{\pi NN}$ of 1250~MeV/c. This value corresponds with those that are typically obtained in the latest parametrizations of the Bonn potential. In the above expression, $G_{\Delta}^{I}$ ($G_{\Delta}^{II}$) denotes the propagator for a $\Delta$ resonance that is created after (before) photoabsorption. In the calculations we adopt the following propagators \begin{eqnarray} G_{\Delta}^{I} & = & {{1} \over {-\sqrt{s_{\Delta}^I} + M_{\Delta} - {i \over 2} \Gamma _{\Delta}^{res}+ V_{\Delta}}} \nonumber \\ \label{eq:delta1} G_{\Delta}^{II} & = & {{1} \over {-\sqrt{s_{\Delta}^{II}} + M_{\Delta} }} \; , \label{eq:delta2} \end{eqnarray} where $\sqrt{s_{\Delta}}$ is the intrinsically available energy for the resonance, $M_{\Delta}$=1232~MeV and $\Gamma _{\Delta}^{res}$ the ``free'' $\pi N$ decay width \cite{Oset,dekker} \begin{equation} \Gamma _{\Delta}^{res} = \frac {2} {3} \frac {f _{\pi N \Delta} ^2} {4 \pi} \frac { \mid \vec{p} _{\pi} \mid ^3 } {m _{\pi} ^2} \frac {M_N} {\sqrt{s_{\Delta}}} \; , \end{equation} where $\vec{p} _{\pi}$ is the decay momentum in the center-of-mass (c.o.m) frame of the $\pi N$ system. As we are dealing with off-shell nucleons the photon energy is not completely available for internal excitation of the $\Delta_{33}$ resonance. This effect is partially responsible for the observed (real) energy shift of the $\Delta$ in the medium. A reasonable substitution for $s_{\Delta}$ is \cite{thomas,reply} \begin{eqnarray} \left( s_{\Delta}^I \right) ^2 & = & -q_\mu q^{\mu} + \left(M_N - \epsilon _h \right)^2 + 2 \omega (M_N - \epsilon _h) \\ \left( s_{\Delta}^{II} \right)^2 & = & \left( \sqrt{ M_N^2 + \mid \vec q \mid ^2 } - \omega \right)^2 \; , \end{eqnarray} where $\epsilon _h$ is the binding energy of the mean-field orbit on which the pion is reabsorbed. The above expression accounts for the observation that the delta peak in ($e,e'$) spectra section shifts to higher $\omega$'s with increasing momentum transfer. Various pion-nucleus \cite{osterfeld}, real photoabsorption \cite{bianchi} and electron scattering studies \cite{koch,oconnell} experiments have pointed towards to strong medium modifications of the $\Delta _{33}$ resonance. For that reason, a medium-dependent term $V_{\Delta}$ that accounts for the interaction of the $\Delta$ resonance with the nucleus was added to the propagator written in (Eq.~\ref{eq:delta1}). Various models that implement a dynamical description of isobar propagation in the medium have been developed \cite{koch,osterfeld}. These models are generally reasonably successful in describing the cross sections for inclusive ($e,e'$) reactions in the $\Delta _{33}$ resonance region \cite{oconnell,cenni84}. The medium modifications of the $\Delta$-resonance appear to be better under control in electromagnetically induced processes than for example in pion-absorption reactions \cite{korfgen94}. For example, Chen and Lee \cite{chen} have shown that a fairly good description of the $^{12}$C($e,e'$) cross sections could be achieved with a $\Delta$ propagator of the type as written in Eq.~(\ref{eq:delta1}) provided that one introduces a simple medium correction of the type $V_{\Delta} [\mathrm{MeV}] = -30-40 \; \mathrm{i}$. Similar values for the $\Delta$-mass shift and broadening have been found in other theoretical approaches \cite{kondra94} and were deduced from recent total photoabsorption measurements \cite{bianchi}. For the calculations presented here, we have used the $\Delta$-medium potential $V_{\Delta}$ as it was calculated by Oset and Salcedo \cite{oset87}. In Figure \ref{fig:width} the imaginary part of $V_{\Delta}$ is shown. In the resonance region, this calculation does indeed reproduce the earlier quoted value for the broadening of the resonance. The above two-body current operators have been used to calculate $^{12}$C($\gamma,pp$) and $^{12}$C($\gamma,pn$) total cross sections in the resonance region. Not only could the position and width of the resonance be reasonably reproduced, also the ($\gamma,pp$)/($\gamma,pn$) ratio for the various shell-model orbits agreed fairly well with experiment \cite{douglas98}. These results lend confidence in the model assumptions with respect to the $\Delta$ currents and propagators. Introducing the earlier discussed two-body current operators in exclusive single-nucleon knockout calculations one ends with the diagrams sketched in Figure~\ref{fig:diagram}. The diagrams of Figure \ref{fig:diagram}(b),(c) and (d) are the pion-exchange contributions to single-nucleon knockout and necessarily imply a charge-exchange mechanism carried by a charged pion. As a consequence, for proton knockout the sum over all occupied single-particle states in the target nucleus ($\sum _{h'}$) produces solely non-zero contributions for the neutron states. The diagrams from Figure~\ref{fig:diagram}(e)-(h) involve $\Delta _{33}$ excitation after interaction of one of the target's nucleons with the (virtual) photon field. For this class of diagrams both charged and neutral pion exchange belongs to the possibilities. The contributions drawn in Figures~\ref{fig:diagram}(e) and (h) involve exclusively neutral pion exchange. Finally, the type of processes drawn in Figure~\ref{fig:diagram}(i)-(l) involve a pre-formed $\Delta$ that is deexcited after interacting with the photon field. Technically speaking, the introduction of the two-body current operators implies that for each combination of a multipole component of the electromagnetic transition operator $T_{J}$ and a partial wave of the ejectile's wave function ($\mid p(lj\epsilon) >$), two-body matrix elements of the type \begin{eqnarray} \label{eq:two} & & <p(lj\epsilon) h^{-1} ; \; J \mid\mid T_{J}^{[2]}(q)\mid\mid 0^{+}(g.s.)> = \sum_{h' J_{1}J_{2}} \sqrt{2J_{1}+1}\sqrt{2J_{2}+1}(-1)^{j_{h}-j_{h'} -J-J_{2}} \left\{ \begin{array}{ccc}j_{h} &j_{h'}& J_{1}\\J_{2} &J& j \end{array} \right\}\nonumber\\ &&\times \Bigl( <h h' \; ; \; J_{1} \mid\mid T_{J}^{[2]}(q)\mid\mid p(lj \epsilon) h' \; ; \; J_{2}> - (-1)^{j'_h+j+J_2} <h h' \; ; \; J_{1} \mid\mid T_{J}^{[2]}(q)\mid\mid h' p(lj \epsilon) \; ; \; J_{2}> \Bigr) \; , \label{eq:mattwo} \end{eqnarray} are to be coherently added to the conventional one-body current contribution from the Impulse Approximation (IA) \begin{equation} <p(lj\epsilon) h^{-1} ; \; J \parallel T_{J}^{[1]}(q) \parallel 0^{+}(g.s.)> \; . \end{equation} In the above expression, $T_{J}^{[1]}$ and $T_{J}^{[2]}$ is the one and two-body current contribution to the electromagnetic transition operator. The explicit expressions for the reduced two-body matrix elements with the meson-exchange and isobar currents can be found in Refs.~\cite{jan97,jana568}. The sum over $h'$ involves all occupied proton and neutron single-particle states. It speaks for itself that for a nucleus like $^{208}$Pb this summation can only be performed at a large computational cost. For consistency reasons and to avoid orthogonality deficiencies, the wave functions for all occupied states $h'$ are calculated in exactly the same mean-field potential in which also the distorted outgoing nucleon wave and the overlap wave function $\left<J_R M_R (E_x) \mid J_i M_i \right>$ is determined. We remark that the very same type of matrix elements that dictate the two-body current contributions to the single-nucleon knockout processes, determine the two-body current contributions to the cross sections for the A($\gamma,NN$) and A($e,e'NN$) processes that are presently the subject of investigation at various laboratories. It speaks for itself that two-nucleon knockout processes represent an intrinsically superior way of exploring the two-nucleon effects in nuclei. The model assumptions with respect to the current operators adopted here, are indentical to those that we have adopted in our two-nucleon knockout studies. Also the potential in which the bound and scattering states are calculated are indentical for the single- and two-nucleon knockout studies. To illustrate the potential of two-nucleon knockout studies to acquire a precise understanding of two-nucleon mechanisms in finite systems, Figure~\ref{fig:gercoeep} shows a comparison of recently obtained $^{16}$O($e,e'pp$) data and our model calculations for the three lowest bins in the excitation energy spectrum in $^{14}$C. The comparison between the calculations and the data is done as a function of the pair missing momentum $\mid \vec{P} \mid = \mid \vec{p}_1 + \vec{p}_2 - \vec{q} \mid$, which is the c.m. momentum of the pair before it undergoes the electromagnetic interaction with the photon field and is established to be the scaling variable in two-nucleon knockout processes, i.e. the counterpart of the variable $\mid \vec{p}_m \mid = \mid \vec{p}_N - \vec{q} \mid$ in the ($e,e'p$). For the data shown in Figure~\ref{fig:gercoeep} the experimental resolution in the excitation (or missing) energy of $^{14}$C was of the order of 4 MeV and the individual states could not be resolved in the missing energy spectrum. On the other hand, high-resolution $^{16}$O($e,e'pp$) investigations from Mainz \cite{guenther} and $^{15}$N($d,^3He$)$^{14}$C transfer reactions \cite{kaschl71} allow to infer that the lowest excitation-energy bin (-4~MeV $\leq E_x \leq $4~MeV) is exclusively fed through the $^{14}$C ground-state transition, whereas the second (4~MeV $\leq E_x \leq $9~MeV) and third bin (9~MeV $\leq E_x \leq $14~MeV) are mainly fed through the $\left| 2^+ ; E_{x} = 7.01,8.32~\mathrm{MeV} \right>$ doublet and the $\left| 1^+ ; E_{x}=11.3~\mathrm{MeV} \right>$ state respectively. These states can be safely quoted to be the only ones in the low excitation-energy spectrum of $^{14}$C that have a dominant ``two-hole'' structure relative to the ground state of $^{16}$O. This information allows to constrain the nuclear-structure input that is required for the $(e,e'pp)$ calculations. The overall agreement between the calculations, that are parameter-free, and the data is reasonable and lends support for the model assumptions with respect to the $\Delta$-current operator and propagator. Indeed, for the transitions to the 2$^+$ and $1^+$ states intermediate $\Delta$ creation is predicted to be the dominant reaction process in the ($e,e'pp$) reaction under evaluation. The shaded region in Figure~\ref{fig:gercoeep} is the calculated contribution from the central Jastrow correlations to the two-proton knockout cross sections. They make a big contribution to the ground-state transition (upper panel) in the low missing momentum regime $P \leq 250$~MeV/c which can be inferred to arise from diproton knockout from heavily correlated $^{1}S_0$(T=0) pairs \cite{gercoprl2}. For the results presented in Figure~\ref{fig:gercoeep} the contribution from the Jastrow or short-range correlations (SRC) was calculated with a correlation function as it was obtained from a G-matrix calculation by Gearhart and Dickhoff \cite{gearhart}. Recently, this correlation function was shown to produce a favorable agreement with $^{12}$C($e,e'pp$) data \cite{blom98}. In comparison with other model predictions for the central correlation function, the one obtained by Gearhart and Dickhoff should be classified in between the categories of ``hard'' (with a core at short internucleon distances) and ``soft'' (characterized by a finite probability to observe nucleon pairs at very short internucleon distances). \section{Results and discussion} \label{sec:results} \subsection{$^{16}$O($\vec{e},e'\vec{p}$) and $^{12}$C($\vec{e},e'\vec{p}$) results for 0.05~$\leq Q^2$~$\leq$0.50~(GeV/c)$^2$} We start our investigations into the role of two-currents for the kinematics of an $^{16}$O($e,e'p$) experiment made at MAINZ and reported in Reference~\cite{blomqvist95}. In this experiment, the cross sections were measured in a broad missing momentum range between 100 and 700~MeV/c. The data were compared to distorted wave impulse approximation (DWIA) calculations. At lower missing momenta ($p_m~\leq$300~MeV/c) the data were observed to overshoot these calculations by a factor of two, whereas for the highest missing momenta probed the opposite effect was noticed. This qualitative behaviour was recently confirmed in an independent calculation by J.J. Kelly \cite{kelly1}. Despite the fact that rather large bound nucleon momenta were probed it was concluded that the deviations from standard (Woods-Saxon) mean-field wave functions are modest. This observation lends support for the picture that nucleon-nucleon correlations are related to the hard-core part of the nucleon-nucleon interaction and are not expected to bring about large deviations from mean-field models in single-knockout processes as long as low excitation energies in the A-1 system are probed. The data of Ref.~\cite{blomqvist95} were obtained for an initial electron energy of 855.1~MeV and $T_p \approx$195~MeV. As a large fraction of the data were taken in parallel kinematics ($\vec{q} \; \| \; \vec{p}_N$), the momentum transfer had to be decreased as higher missing momenta were probed. Consequently, to reach higher missing momenta one had to move out of quasi-elastic kinematics. Figures~\ref{fig:blom1} and \ref{fig:blom2} show the predicted sensitivity of the Mainz results to the two-body currents. The Figures show results for four different central values in $Q^2$ and parallel kinematics. The kinematics is such that one moves out of quasi-elastic conditions as higher missing momenta are probed. As is commonly done, we present the angular cross-sections in terms of the reduced cross section ``$\rho _m$''. Hereby, we adopt the standard convention of dividing the calculated cross sections by a kinematical factor times the so-called ``CC1'' \cite{forest} elementary electron-proton cross section \begin{equation} \rho_m \equiv \frac {\frac {d^5 \sigma } {d \Omega _N d \epsilon ' d \Omega _{\epsilon '}}} {{\frac {p_N E_N} {(2\pi)^3}} f_{rec}^{-1} \sigma_{ep}^{CC1}} \; . \end{equation} Referring to Figures \ref{fig:blom1} and \ref{fig:blom2}, striking features are: (1) the sensitivity of the cross sections to the two-body currents is substantial and (2) the two-body currents do not dramatically alter the shape of the reduced cross section for none of the four $Q^2$ values considered. As a consequence, the impact of the two-body currents would generally not be noticed when comparing results of calculations performed within the IA with data but would simply be ``effectively'' accounted for in the spectroscopic factor that is used to scale the DWIA calculations to the data. The net effect of the two-body currents is a reduction of the cross sections. This reduction is of the order 10-20\% for the lowest missing momenta range and can amount to a factor of two at $p_m \approx 300$~MeV/c. Note that for the highest missing momenta considered in Figures \ref{fig:blom1} and \ref{fig:blom2} inclusion of the two-body currents tends to increase the cross section. This is generally the case when higher missing momenta are probed \cite{kester96}. The observed reduction at low missing momenta that was ascribed to two-body current effects is not sufficient to explain why the data are substantially lower (factor of two) than expected on the basis of DWIA models. Indeed, the curves of Figures \ref{fig:blom1} and \ref{fig:blom2} are obtained with a reduction factor of 0.25 for the ground-state transition and 0.30 for the $^{16}$O($e,e'p$)$^{15}$N(3/2$^-$,6.32~MeV) transition. With these values we obtain reasonable visual fits of the data. The transparency at the considered four-momentum transfers $Q^2$ is experimentally determined to be about 0.75 for a light nucleus like $^{12}$C \cite{potterfeld,garino92}. When correcting the reduction factors we have applied to obtain visual fits of the data, for transparency (or attenuation) corrections that fall beyond the scope of our model calculations, we obtain the spectroscopic factors $S_{lj} (1/2^-,$E$_x$=0.0~MeV)$\approx$ 0.67 and $S_{lj}(3/2^-,$E$_x$=6.3~MeV)$\approx$1.6. In line with the theoretical conclusions drawn from the analyses of References~\cite{blomqvist95} and \cite{kelly2} these values are substantially smaller than the values obtained with high-resolution measurements at lower proton kinetic energies where it was found that $S_{lj} (1/2^-,$E$_x$=0.0~MeV)=1.20 and $S_{lj}(3/2^-,$E$_x$=6.3~MeV)=2.00 \cite{leuschne94}. Note, however, that for $1p_{3/2}$ knockout the spectroscopic factor that is deduced from the calculations that include the two-body currents is about 30\% larger than the value that would be deduced in the IA. Another striking feature of the results displayed in Figures \ref{fig:blom1} and \ref{fig:blom2} is that the influence of the subnuclear degrees-of-freedom is more pronounced for knockout from the ``stretched'' ($j_h= l_h + \frac{1}{2}$) orbit (i.e. $1p_{3/2}$) than from its ``jack-knifed'' spin-orbit partner ($j_h= l_h - \frac{1}{2}$) (i.e. $1p_{1/2}$). The major reason for this behaviour is simply that all the two-body current operators have a strong spin dependency. It is worth mentioning that the relativistic effects arising from the lower (or negative energy) components in the bound state wave functions were recently shown to have a similar sort of sensitivity. Indeed, the (relativistic) $(e,e'p)$ results for knockout from the ``stretched'' orbits will be closer to their respective non-relativistic limits than those for ``jack-knifed'' orbits \cite{caba98}. Accordingly, the predicted relativistic effects attributed to the small components in the bound state wave functions are smallest for these transitions with maximal two-body current contributions. The question arises whether the relatively strong sensitivity to two-body currents noticed in Figures~\ref{fig:blom1} and \ref{fig:blom2} is an intrinsic property of the specific $Q^2$ range probed or is rather due to the fact that these results were acquired at smaller values of x and subsequently clearly out of quasi-elastic conditions. Results for real quasi-elastic kinematics at about the same value for the proton's kinetic energy and initial electron energy are shown in Figures~\ref{fig:mommainz}. These results are obtained in quasi-perpendicular kinematics. Here, the variation in missing momentum is reached by varying the polar angle $\theta _p$ of the ejected proton. For the remainder of this paper all calculated (reduced) cross sections are shown for full sub-shell occupancy, i.e. $S_{lj}=(2j+1)$. It becomes clear that under strict quasi-elastic conditions the effect of the subnuclear degrees-of-freedom is substantially smaller than at smaller values of Bjorken x. It is worth mentioning that the results of Figure~\ref{fig:mommainz} are obtained in more transverse kinematics than those of Figures~\ref{fig:blom1} and \ref{fig:blom2}. Accordingly, with the eye on minimizing the effect of two-body currents it is very essential to move in quasi-elastic conditions {\sl even if this implies shifting to kinematics which is conceived to be transverse.} Another observation is that the extent to which to the two-body currents are important does not depend on the polar angle $\theta _N$ of the ejected nucleon. Figures \ref{fig:strmainz} and ~\ref{fig:polmainz} show the structure functions and polarization observables for the same kinematics in which the cross sections of Figure~\ref{fig:mommainz} are obtained. In presenting the structure functions we prefer not to divide out any kinematical variables from the different contributions to the total angular cross section. Formally, this means that we write Eq.~(\ref{eq:eep}) in the form \begin{equation} {d^5 \sigma \over d \Omega _N d \epsilon ' d \Omega _{\epsilon '}} \equiv \sigma_L + \sigma_T + \sigma_{LT} + \sigma_{TT} +h \left(\sigma ' _{LT} + \sigma ' _{TT} \right) \; , \end{equation} where the precise definition of all contributing terms is obvious. This way of presenting results has the outspoken advantage that the relative contribution of each term in the cross section can be easily evaluated. Furthermore, possible confusions regarding the adopted definitions for the different structure functions are avoided. In line with the observations made for $^{4}$He and the deuteron \cite{gilad98} case, the interference terms that are an order of magnitude smaller than $\sigma _L$ and $\sigma _T$ exhibit the strongest sensitivity to contributions that go beyond the IA. We now turn to the discussion of the recoil polarization observables. For coplanar kinematics the sole non-vanishing recoil polarization observables are $P_n,P'_l$ and $P'_t$. The $P_n$ does vanish identically in the plane-wave impulse approximation. As such it allows to constrain the model assumptions with respect to the final-state interaction (FSI). Inspecting Figure~\ref{fig:polmainz} it is clear that at low missing momenta the qualitative behaviour of the induced polarization $P_n$ turns out to be relatively insensitive to the current contributions beyond the impulse approximation. This result confirms the findings of the Pavia calculations reported in Ref.~\cite{mandevil94}. The first data set for the induced polarization in finite nuclei has recently become available \cite{woo98}. In Figure~\ref{fig:woo} the calculated reduced cross section and recoil polarization observables are shown for the kinematics of this experiment. For knockout from the $1s$ shell, where the nucleon is mainly located at high nuclear densities, the influence of the subnuclear d.o.f. is more pronounced than for $p$-shell knockout. Given that we do not rely on any empirical input for the description of the FSI, the agreement with the data is satisfactory but inferior to the quality of agreement that was reached with the calculations by J.J.~Kelly \cite{woo98}. These calculations constrain the description of the final-state interactions with the aid of optical potentials derived from the analysis of hadronic induced reactions. For most observables, the meson-exchange and isobar effects are acting in opposite directions. At low missing momenta the induced polarization exhibits a small sensitivity to the two-body currents. At moderate and high missing momenta the $P_n$ is predicted to exhibit a large sensitivity to two-body currents. A similar observation is made for the polarization transfer observables. Recently it was shown that the recoil proton polarizations for p($\vec{e},e'\vec{p}$) and d($\vec{e},e'\vec{p}$)n are almost identical \cite{milbrath98}. From the theory side, rather small contributions from the MEC and IC in the double polarization observables were predicted in the deuteron \cite{gilad98} and $^{4}$He \cite{laget94} case in quasi-elastic kinematics. Despite the fact that the polarization transfer variable $P'_l$ and $P'_t$ are related to the interference structure functions $W'_{TT}$ and $W'_{LT}$ their predicted sensitivity to the two-body currents looks relatively small. Recently, the polarization transfer observables raised considerable interest in that they would provide a direct handle on the (possible) medium dependence of the nucleon form factors. Indeed, in the plane-wave impulse approximation (PWIA) it can be shown that for electroexcitation from a free proton \begin{equation} \frac {P_l'}{P_t'} = - \frac {G_M^p} {G_E^p} \frac {(\epsilon + \epsilon ') tan \frac {\theta_e} {2}} {2 M_p} \; . \end{equation} It is of the utmost importance to investigate all possible mechanisms that could bring about changes in the above ratio without being related to medium modifications of the form factors. Whereas it was recently shown that final-state interaction and gauge ambiguities effects are only marginally affecting the ratio $\frac {P_l'}{P_t'}$ \cite{kelly} the question arises whether meson-exchange and isobaric currents could bring about any change in the ratio of the double polarization observables. Whereas the impact of the subnuclear d.o.f. on the value of polarization observables looks small in quasi-elastic kinematics and small missing momenta, the MEC and IC are not necessarily unimportant, especially for such delicate issues like nucleon form factor studies where the medium dependency is predicted to be modest \cite{lu98b} and the projected accuracy of the planned experiments is astounding. With the aim of studying the sensitivity of the ratio $\frac {P_l'}{P_t'}$ to meson-exchange currents and intermediate isobar creation we have calculated the following double ratio \begin{equation} \frac { \left( \frac {P_l'}{P_t'} \right)} { \left( \frac {P_l'}{P_t'} \right)^{IA}} \end{equation} where ``$IA$'' refers to the calculated ratio in the impulse approximation, this is when retaining solely the one-body currents in the calculations but keeping all other ingredients fixed. As could be expected, the observed deviation from the IA result enhances as the missing momentum increases. The most favorable regime to perform the recoil polarization measurements for knockout from p-shell orbits, is probably in the peak of the cross section ($p_m \approx$100~MeV/c). For the $^{16}$O($e,e'p$)$^{15}$N(1/2$^-$) transition, the deviation in the ratio $\frac {P_l'}{P_t'}$ that is attributed to the MEC and IC is about 5-10\% in the peak of the cross section. For knockout from its spin-orbit partner ($1p_{3/2}$) the calculated effect is 10-15\%. In both cases, at 200~MeV/c the modifications in the ratio due to two-body currents has grown to 20\% and with increasing $p_m$ the effect appears to be rather out of control. \subsection{$^{16}$O($\vec{e},e'\vec{p}$) results at $Q^2$=0.8~(GeV/c)$^2$ and x$\approx$1} We continue our investigations into the role of two-body currents by considering the kinematics of the TJNAF experiments E89-003 and E8-9033 \cite{e89033}. The E89-003 experiment has measured the separated structure functions and momentum distributions for $^{16}$O($e,e'p$) under quasi-elastic conditions at $\epsilon$=2.445~GeV, $\omega$=445~MeV and q=1~GeV/c. In E89-033 on the other hand, the $^{16}$O($\vec{e},e' \vec{p}$) polarization observables were measured for approximately the same kinematics. Figures \ref{fig:mome89003}-\ref{fig:sstate} summarize the calculated results for the reduced cross sections, structure functions and recoil polarization observables for knockout from the two $p$-shell and the $s$-shell state. In comparing these results with those obtained in the previous Subsection for quasi-elastic conditions at lower four-momentum transfer, we can study the $Q^2$ dependency of the meson and isobar degrees of freedom. In analogy with the obvervations made at lower $Q^2$, the pion-exchange currents (``MEC'') increase the cross section and this effect is completely counterbalanced by the $\Delta$ current that produces the largest contributions. The $Q^2$ dependency of the two-body current effects can maybe be better estimated from the individual structure functions. Comparing Figures~\ref{fig:strmainz} and \ref{fig:str3e89003} one does indeed observe a decreasing trend in the relative importance of the two-body currents. A similar tendency is observed for the recoil polarization observables (Figure~\ref{fig:pol2a3}. The deviation in the ratio $\frac {P_l'}{P_t'}$ in the peak of the cross section is of the order of a few percent for the ground-state transition (knockout from the $1p_{1/2}$ orbit), whereas for the excitation of the $3/2^-$ hole state at 6.32~MeV excitation energy in $^{15}$N (knockout from the stretched $1p_{3/2}$ orbit) the effect is close to 10\%. With the aim of studying the medium-dependent effects one could be tempted to probe the region of highest nuclear density which for a nucleus like $^{16}$O implies studying the region of $1s_{1/2}$ knockout. In light of the mass-independence of the two-body current effects that is frequently alluded to, our findings for $1s_{1/2}$-knockout in $^{16}$O can to a certain extent serve as a guideline for the effects than could be expected in $^4$He. From all orbits studied here, knockout from the $1s_{1/2}$ orbit exhibits the strongest sensitivity to the two-body currents. After all, this observation is not that surprising given its stretched status and the fact that the region of highest nuclear density is probed. For example, the $LT$ and $TT$ interference responses in Figure~\ref{fig:str1e89003} grow substantially after including the two-body currents. Indications for the strong sensitivity of the ${LT}$ structure function to two-body currents was reported for the $^{4}$He case in Ref.~\cite{epstein93}. Also for the recoil polarization observables in the missing-energy range of $1s_{1/2}$ knockout (Figure~\ref{fig:sstate}), sizeable contributions from the meson-exchange and isobar currents are predicted. For the ratio $\frac {P_l'}{P_t'}$ the two-body currents bring about a reduction which is slightly bigger than 10\%. This is qualitatively very similar with the result obtained for the ``stretched'' p-shell state ($1p_{3/2}$) (Figure~\ref{fig:pol2a3}) so that some general qualitative behaviour for the sensitivity of the $\frac {P_l'}{P_t'}$ ratio to multi-nucleon currents seems to emerge. Indeed, at low missing momenta and quasi-elastic conditions the two-body currents reduce the ratio $\frac {P_l'}{P_t'}$, an effect that slowly decreases as one goes to higher $Q^2$ and is larger for knockout from ``stretched'' than from the ``jack-knifed'' single-particle orbits. Finally we turn our attention to the role of the relativistic corrections (Section~\ref{sec:current}) in the one-body current operator. The dotted lines in Figures~\ref{fig:str2e89003}, \ref{fig:str3e89003} and \ref{fig:str1e89003} show the calculated contributions to the cross section when neglecting the relativistic corrections in the one-body current operator, i.e. when adopting $C_1(q,Q^2)=C_3(q,Q^2)=1$ and $C_2(q,Q^2)=0$ in Eq.~(\ref{eq:redcurrent}). When comparing these curves with the solid ones one can estimate the effect of the relativistic corrections in the one-body current. For the transverse responses $\sigma _T$ and $\sigma _{TT}$ the Foldy-Wouthuysen prescription does not lead to any relativistic correction into lowest order. Whereas the spin-orbit term in the charge-density operator produces rather small corrections in the longitudinal response $\sigma _L$ in the considered kinematics, the longitudinal-transverse response $\sigma _{LT}$ for the ``stretched'' single-particle states $1p_{3/2}$ and $1s_{1/2}$ is doubled after including the relativistic corrections in the charge-density operator. A strong sensitivity of the $LT$ response to relativistic corrections was earlier found for $d(e,e'p)$ \cite{gilad98,hummel94,ducret94}. \subsection{Neutron knockout} To our knowledge, beyond the giant resonance region no ($e,e'n$) measurements have been made. Nevertheless, neutron knockout investigations with moderate energy resolution could be done at MAMI, BATES or TJNAF. Results for neutron knockout in kinematic conditions that can be reached with electron beam energies below 1~GeV are shown in Figure~\ref{fig:een}. These results are obtained with a neutron formfactor $G_E^n$=0. The $(e,e'n)$ calculations were done in the same kinematical conditions for the ($e,e'p$) curves of Figure~\ref{fig:mommainz}. The reduced cross sections for neutron knockout are subject to larger two-body current corrections than corresponding proton results. As a matter of fact, the absolute magnitude of the two-body current contributions is comparable for proton and neutron knockout. The fact that $\left( G_M^p/ G_M^n \right)^2 \approx$2 and the absence of a substantial neutron charge, however, makes them to be relatively more important in the neutron knockout channel. As could be expected on the basis of the isospin structure of the underlying current operators, the qualitative behaviour is similar for proton and neutron knockout. The net reduction of the cross sections displayed in Figure~\ref{fig:een} is of the order 10-20\%. Note that coupled-channel-effects between the proton knockout channels and the considerably weaker ($e,e'n$) channels are expected to compensate largely for the loss of strength brought about by the meson and isobar degrees-of-freedom \cite{kelly2,jan89}. Our ($e,e'n$) results are qualitatively similar to those reported in \cite{boffi92}. The MEC are observed to increase the cross section, an effect that is largely overcompensated by the isobaric current that works in the opposite direction. This reduction can be attributed to the fact that the one-body magnetization current and the isobaric current interfere destructively. \subsection{$^{208}$Pb($e,e'p$) results at $Q^2$=0.3~(GeV/c)$^2$ and x$\approx$1} In this section we report on calculations for an $^{208}$Pb($e,e'p$) experiment that was done at NIKHEF. This experiment was performed in very transverse kinematics ($\theta_e=$96$^o$) at an incident electron energy of 462.14~MeV and near quasi-elastic conditions ($T_p$=161.~MeV, x=0.94). As this corresponds with very ``transverse'' kinematics we consider this as a worst-case-scenario benchmark for investigating the sensitivity to two-currents that could be expected in quasi-elastic ($e,e'p$) from heavy nuclei. This is particularly important in view of the fact that the $^{208}$Pb target is sometimes advocated as an appropriate surrogate for nuclear matter studies. In comparison with studies on light nuclei, for heavy nuclei there is a price to pay in that Coulomb distortion effects and reduced nucleon transparencies make both the electromagnetic and the FSI part of the reaction process more difficult to handle and subject to likely enhanced theoretical uncertainties. Here, we only want to study the {\em relative contribution} from the two-body currents that could be expected for a heavy nucleus like $^{208}$Pb. For computational reasons we have neglected the Coulomb distortion effects. Indeed, an exact treatment of these would imply that the two-body matrix elements of Eq.~(\ref{eq:mattwo}) that involve a sum over all nucleons in the target nucleus, are to be calculated for a whole range of q values. Whereas calculations of this type could be performed at lower momentum transfer \cite{veerle2}, the amount of multipoles required at higher momentum transfer makes it an enormous computational task. Figure~\ref{fig:pb208} displays the reduced $^{208}$Pb($e,e'p$) cross section for knockout from the $3s_{1/2}$ and $2d_{5/2}$ orbits. It becomes clear that the predicted two-nucleon effects are modest and of the same relative size in comparison with the findings for light nuclei. Whereas the transparency for nucleon knockout has been established to decrease as the mass number increases \cite{potterfeld} the impact of the two-body currents on single-nucleon knockout appears to be rather mass independent. \section{Summary and Conclusions} \label{sec:conclusions} Within the context of ongoing research activities in two-nucleon and pion production from nuclei a fair understanding of the two-nucleon currents in finite nuclei has been reached. With this knowledge the contribution of two-body currents to exclusive ($e,e'p$) and ($e,e'n$) reactions can be calculated with some confidence. Reasonable estimates of their contribution is essential in view of the fact that subnuclear d.o.f. frequently represent unwanted background that could be at the origin of ambiguities in the interpretation of ($\vec{e},e'\vec{p}$) studies. Most of our investigations were carried out in quasi-elastic kinematics and momentum transfers $\mid \vec{q} \mid \le$~1~GeV/c. The effect of two-body currents dramatically increases as one moves out of quasi-elastic conditions and/or higher missing momenta are probed. In quasi-elastic kinematics and the whole range of $Q^2$ values studied here, the two-body currents decrease the $(e,e'p)$ differential cross sections when lower missing momenta are probed. As the shape does not seem to be substantially affected by the two-body currents it appears virtually impossible to acquire experimental evidence for the role of two-body current effects from plain $(e,e'p)$ cross-section measurements in quasi-elastic kinematics. At the same time, our findings imply that the spectroscopic factors as they are usually derived from IA calculations are subject to corrections that would make them bigger. These corrections are not unimportant and exhibit a $Q^2$ dependence. Even at four-momentum transfers of $Q^2 \approx$~0.8~(GeV/c)$^2$ the effect of the two-body currents on the spectroscopic factor can be as large as 20\%. The role of the two-body currents can be made more explicit by measuring the interference response functions that, admittedly, represent a rather small part in the total angular cross section. The induced polarization $P_n$ exhibits a moderate sensitivity to the two-body currents at low missing momenta $p_m \le 200$~MeV/c. Therefore, it retains its status as a proper variable to constrain the final-state interaction effects. Similar sort of sensitivities are found for the polarization transfer variables. In the light of exploiting these variables to increase our understanding of the (possible) medium dependency in the nucleon form factors, the effect of the meson and isobar degrees of freedom on the ratio $\frac {P_l'}{P_t'}$ can be boiled down to a few percent at higher values of $Q^2$. Under typical MAMI kinematics, which implies four-momentum transfers of the order $Q^2 \le $0.50~(GeV/c)$^2$, the impact of subnuclear degrees of freedom on the $\frac {P_l'}{P_t'}$ ratio is predicted to be of the order of 10\%. Of all pion-related two-body currents studied here the $\Delta$-isobar current has by far the strongest impact on the ($\vec{e},e'\vec{p}$) and ($e,e'n$) observables. Multi-coincidence experiments of the type ($\gamma^{(*)},NN$) and ($\gamma^{(*)},N\pi$) at higher four-momentum transfers $Q^2$ would greatly help in further constraining the model assumptions with respect to the $\Delta$-isobar current operator when reaching higher energy and momentum transfers. {\bf Acknowledgement} This work was supported by the Fund for Scientific Research of Flanders under Contract No 4.0061.99 and the University Research Council.
\section{Introduction} \label{sec:intro} The field equations of general relativity are usually written in terms of the Einstein tensor and the stress tensor; however there is an alternative expression, called Jordan's formulation of the field equations, in which the field equations are expressed as first order equations in the Weyl tensor \cite{bi:HE}. The Weyl tensor can be expressed, using first order equations, in terms of a three index tensor \cite{bi:lanczos} \cite{bi:mdr88} \cite{bi:mdr89} \cite{bi:mdr92} \cite{bi:mdr94} called the Lanczos tensor. This tensor can be used to produce gravitational energy tensors of the correct dimension \cite{bi:mdr88}, and these can be used to measure the speed of gravitational waves \cite{bi:mdr94}. In Jordan's formulation the field equations are of a similar form to the Maxwell equations in terms of the electro-magnetic field tensor. The electro-magnetic field tensor can be expressed as a first order differential equation in the vector potential, and thus the Lanczos tensor in analogous to the vector potential in electro-magnetic theory. The Lanczos tensor is not the only tensor that can be thought of as being analogous to the vector potential; because the field equations are second order in the metric it is possible also to think of the metric (or the difference between the metric and the Minkowski metric) as being analogous to the the vector potential. There is also the Ashtekar potential in the theory of Ashtekar variables \cite{bi:ashtekar}; this potential is not the same object as the Lanczos tensor, because the equation for the Weyl tensor involves cross terms in the Ashtekar potential, unlike the Lanczos tensor in which the expression for the Weyl tensor is linear. The differential equations involving the Lanczos tensor which govern the dynamics of the field equations do not have a Lagrangian formulation, thus traditional methods of quantization cannot be applied to the field equations in this form. In electro-magnetic theory the vector potential was first introduced in order to express the equations of classical electrodynamics in simpler form. In classical physics the only physical effect of an electro-magnetic field on a charge is the Lorentz force, and this only exists in regions where the electric or magnetic field in non-vanishing. The Aharonov-Bohm effect \cite{bi:ES} \cite{bi:AB} \cite{bi:BH} demonstrates that this is not so in quantum mechanics; physical effects occur in regions where the electric and magnetic fields both vanish, but where the vector potential does not vanish. It has been experimentally confirmed \cite{bi:chambers}. In general relativity the existence of th Lanczos tensor might be just a technical curiosity, or it might have fundamental significance in the way that the vector potential does in electro-magnetic theory. The object of the present paper is to determine a thought experiment, similar to the Aharonov- Bohm experiment, which would in principle determine whether the Lanczos tensor effects the dynamics and so is physically significant. At the quantum level the vector potential enters the Schrodinger equation through the application of the electro-magnetic covariant derivative. The main problem in our approach is what should correspond to this covariant derivative; after all there is already a covariant derivative in general relativity constructed from the Christoffel symbol. Here it is postulated that in the quantum realm a covariant derivative involving the Lanczos tensor plays a role, and that it is the correct covariant derivative to apply to the Schrodinger equation in analogs of the Aharonov-Bohm effect. Now the main problem becomes how precisely should this covariant derivative be constructed; this cannot be know {\it a-priori} and so it is necessary to construct an example which will motivate a suitable definition of a new covariant derivative. If such a covariant derivative could be constructed it is possible to anticipate several difficulties. {\it Firstly} why does the classical theory have no use for a Lanczos covariant derivative? {\it Secondly} in the quantum realm the Christoffel covariant derivative is still necessary to connect the Lanczos tensor to the Weyl tensor, why have two covariant derivatives? {\it Thirdly}, in the Aharonov-Bohm effect the electro-magnetic field is not quantized, only the test particles are treated in a quantum mechanical manner; in our case would only the test particles be treated quantum mechanically, or would the Lanczos tensor or the metric or both be quantum fields? The Aharonov-Bohm effect depends crucially on the existence of a choice of the vector potential which is well-defined and continuous everywhere. For example, as discussed in \cite{bi:BH}, it is possible to choose a gauge in which the vector potential vanishes outside the solenoid and claim that there should be no effect; in fact, as experiments \cite{bi:chambers} vindicate, a gauge should be chosen in which the vector potential is continuous everywhere. It is assumed that a similar criteria on continuity exists for the present examples, so that it is possible to fix the gauge and then calculate the global effect of having a non-vanishing Lanczos tensor in the exterior region of the space-time. In the examples presented here it is found that the Lanczos tensor is either continuous or not depending on whether the derivative of the metric is continuous or not, irrespective of the choice of gauge. Thus there is no criteria to inform us which is the correct gauge, and hence the analogy cannot be carried through completely. This might be because the present examples are so geometrically simple. The Aharonov-Bohm effect depends on a current in a solenoid and it is not clear what is the general relativistic analog of a current in a solenoid. The simple cylindrical space-time used here might just be analogous to a line of charges, for which the Aharnov-Bohm effect does not work. Perhaps the correct analogy is a fluid in a pipe, i.e. a cylindrical space-time with a perfect fluid moving along the axis; however such an exact solution is not known. The example discussed here use a simple cylindrically symmetric space-time. This space-time is general enough to include the simplest cosmic string \cite{bi:vilenkin} \cite{bi:gott}. The approach used here relies on the Aharonov-Bohm effect being a quantum mechanical effect, and should not be confused with classical analogs of the Aharonov-Bohm effect which exist in cosmic string and some other space-times \cite{bi:bezerra}. In section \ref{sec:II} the elementary properties of the Lanczos tensor are expounded. In section \ref{sec:III} the Lanczos tensor is produced for a simple cylindrically symmetric space-time. In section \ref{sec:IV} the construction of covariant derivatives involving the Lanczos tensor are discussed. \section{The Lanczos Tensor.} \label{sec:II} The field equations of general relativity can be re-written in Jordan's form \cite{bi:HE} \begin{eqnarray} &&C^{~~~d}_{abc.;d}=J_{abc},\nonumber\\ &&J_{abc}=R_{ca;b}-R_{cb;a}+\frac{1}{6}g_{cb}R_{;a}-\frac{1}{6}g_{ca}R_{;b}, \label{eq:1} \end{eqnarray} which is analogous to Maxwell's equations \begin{equation} F^{~b}_{a.;b}=J_a. \label{eq:2} \end{equation}} %\indent The Weyl tensor can be expressed in terms of the Lanczos tensor \cite{bi:lanczos} \cite{bi:mdr88} \cite{bi:mdr89} \cite{bi:mdr92} \begin{eqnarray} C_{abcd}&=&H_{abc;d}-H_{abd;c}+H_{cda;b}-H_{cdb;a}\nonumber\\ &-&(g_{ac}(H_{bd}+H_{db})-g_{ad}(H_{bc}+H_{cb})+ g_{bd}(H_{ac}+H_{ca})-g_{bc}(H_{ad}+H_{da}))/2\nonumber\\ &+&2H^{ef}_{..e;f}(g_{ac}g_{bd}-g_{ad}g_{bc})/3, \label{eq:3} \end{eqnarray} where the Lanczos tensor has the symmetries \begin{equation} H_{abc}+H_{bac}=0,\\ \label{eq:4} \end{equation}} %\indent \begin{equation} H_{abc}+H_{bca}+H_{cab}=0, \label{eq:5} \end{equation}} %\indent and where $H_{bd}$ is defined by \begin{equation} H_{bd}\equiv H^{~e}_{b.d;e}-H^{~e}_{b.e;d}. \label{eq:6} \end{equation}} %\indent Equation \ref{eq:3} is invariant under the algebraic gauge transformation \begin{equation} H_{abc}\rightarrow H'_{abc}=H_{abc}+\chi_a g_{bc}-\chi_b g_{ac}, \label{eq:7} \end{equation}} %\indent where $\chi_a$ is an arbitrary four vector. The Lanczos tensor with the above symmetries has 20 degrees of freedom, but the Weyl tensor has 10. Lanczos \cite{bi:lanczos} reduced the degrees of freedom to 10 by choosing the Lanczos algebraic gauge \begin{equation} 3\chi_a=H^{~b}_{a.b}=0, \label{eq:8} \end{equation}} %\indent and the Lanczos differential gauge \begin{equation} L_{ab}=H^{~~c}_{ab.;c}=0. \label{eq:9} \end{equation}} %\indent These gauge choices are in some ways different than those in electro-magnetic theory. The algebraic gauge is different because it is algebraic and not differential in nature. The differential gauge is different because a differential gauge transformation alters components which do not participate in constructing the Weyl tensor; in electro-magnetic theory a gauge transformation alters components in the vector potential all of which participate in constructing the electro-magnetic tensor. These difference are well illustrated by the example in the next section. When the Lanczos tensor happens to be the gradient of an anti-symmetric tensor of the second order \begin{equation} H_{abc}=F_{ab;c}, \label{eq:10} \end{equation}} %\indent and if the Lanczos tensor is in the algebraic gauge \ref{eq:8}, \ref{eq:5}, and \ref{eq:10} imply that $F_{ab}$ obeys Maxwell's equations. It is not possible to introduce a source $J_a$ to \ref{eq:10} without it having an un-natural constraint by virtue of the identity \ref{eq:3}. In the case of weak gravity \begin{equation} g_{ab}=\eta_{ab}+h_{ab}, \label{eq:11} \end{equation}} %\indent where $\eta_{ab}$ is the Minkowski metric and $h_{ab}$ and its derivatives are small, the Lanczos tensor can be written as \begin{equation} 4H_{abc}=\partial_b h_{ac}-\partial_a h_{bc} +\frac{1}{6}h_{,a}\eta_{bc}-\frac{1}{6}h_{,b}\eta_{ac}, \label{eq:12} \end{equation}} %\indent where $h=h^a_{.a}$. \section{The Lanczos Tensor for a Simple Cylindrically Symmetric Space-time.} \label{sec:III} In this section we find the Lanczos tensor for a simple static cylindrically symmetric space-time with line element \begin{equation} ds^2=-dt^2+dr^2+X~d\phi^2+dz^2. \label{eq:13} \end{equation}} %\indent The non-vanishing Christoffel symbols are \begin{equation} \Gamma^r_{\phi\ph}=-\frac{1}{2}X_r,~~~ \Gamma^\phi_{\phi r}=X_r /2X. \label{eq:14} \end{equation}} %\indent The Riemann, Ricci, Einstein, and Weyl tensors are conveniently expressed in terms of the Ricci scalar \begin{equation} R=2R^\phi_{.\phi}=2R^r_{.r} =-2G^t_{.t}=-2G^z_{.z} =2R^\phi_{.r\phi r}=2R^r_{.\phi r\phi}/X, \label{eq:15} \end{equation}} %\indent \begin{equation} R=-6C_{tztz}=6C_{\phi r\phi r}/X=12C_{t\phi t\phi}/X=-12C_{z\phi z\phi}/X =12C_{t\phi t\phi}=-12C_{trtr}, \label{eq:16} \end{equation}} %\indent [note added 1999 which $C_{t \phi t\phi}$ is correct] where \begin{equation} R=-_{rr}X/X+X^2_r/2X=-X^{-\frac{1}{2}}(X'X^{-\frac{1}{2}})'. \label{eq:17} \end{equation}} %\indent This space-time is general enough to include the simple cosmic string for which the metric is \cite{bi:vilenkin} \cite{bi:gott} \begin{equation} ds^2=-dt^2+dr^2+\rho*sin^2(\rho/\rho^*)d\phi^2+dz^2, \label{eq:18} \end{equation}} %\indent where $\rho^*=(8\pi\epsilon)^\frac{1}{2}$ and $\epsilon$ is the density of the string. The Ricci scalar is \begin{equation} R=4\pi\epsilon. \label{eq:19} \end{equation}} %\indent At the join $r=r_0$, $\rho=\rho_0$ the interior metric is attached to the exterior metric \begin{equation} ds^2=-dt^2+dr^2+a^2r^2d\phi^2+dz^2, \label{eq:20} \end{equation}} %\indent where $a$ is given by \begin{eqnarray} &&a=1-4\mu,\nonumber\\ &&\mu=\int^{\rho_0}_0\int^{2\pi}_0\epsilon\rho^*sin(\rho/\rho^*)d\phi d\rho =2\pi\rho^{*2}(1-cos(\rho/\rho^*)), \label{eq:21} \end{eqnarray} and $\mu$ is called the linear energy density. The requirement that the metric is continuous at the join is \begin{equation} ar_0=\rho^* sin(\rho/\rho^*). \label{eq:22} \end{equation}} %\indent The derivative of the metric is continuous at the join, as this is required for there to be no surface stress present; this requirement gives a in \ref{eq:20}, otherwise a would be simply absorbed into the line element. From \ref{eq:3} and \ref{eq:15} we have from the $C_{tztz}$ or $C_{\phi r\phi r}$ component \begin{equation} R=2\partial_rH_{trt}-2\partial_rH_{\phi r\phi}+4\partial_rH_{\phi r\phi}/X +X_rH_{rtr}/X-X_rH_{zrz}/X-2X_rH_{\phi r\phi}/X^2, \label{eq:23} \end{equation}} %\indent from the $C_{t\phi t\phi}$ or $C_{trtr}$ component \begin{equation} R=-4\partial_rH_{trt}-8\partial_rH_{zrz}+4\partial_rH_{\phi r\phi}/X +4X_rH_{trt}/X +2X_rH_{zrz}/X-2X_rH_{\phi r\phi}/X^2, \label{eq:24} \end{equation}} %\indent from the $C_{\phi z\phi}$ or $C_{trtr}$ component \begin{equation} R=8\partial_rH_{trt}+4\partial_rH_{zrz}+4\partial_rH_{\phi r\phi}/X-2X_rH_{trt}/X -4X_rH_{zrz}/X-2X_rH_{\phi r\phi}/X^2, \label{eq:25} \end{equation}} %\indent Subtracting \ref{eq:24} from \ref{eq:23} or \ref{eq:25} from \ref{eq:24} we have \begin{eqnarray} 0&=&2\partial_rH_{trt}+2\partial_rH_{zrz}-X_rH_{trt}/X-X_rH_{zrz}/X\nonumber\\ &=&2\sqrt{X}\partial_r(H_{trt}X^{-\frac{1}{2}})+2\sqrt{X}\partial_r(H_{zrz}X^{-\frac{1}{2}}), \label{eq:26} \end{eqnarray} which integrates to give \begin{equation} H_{trt}=k\sqrt{X}-H_{zrz}, \label{eq:27} \end{equation}} %\indent where $k$ is a constant. From \ref{eq:27} and \ref{eq:23} or \ref{eq:24} or \ref{eq:25} we have \begin{equation} R=2kX_rX^{-\frac{1}{2}}-4\partial_rH_{zrz}+4\partial_rH_{\phi r\phi}/X -2X_rH_{zrz}/X -2X_rH_{\phi r\phi}/X^2. \label{eq:28} \end{equation}} %\indent Here \ref{eq:8} the Lanczos algebraic condition is $-H_{trt}+H_{zrz}+H_{\phi r\phi}/X=0$, it gives \begin{equation} H_{zrz}=k\sqrt{X}/2-H_{\phi r\phi}/2X. \label{eq:29} \end{equation}} %\indent Equations \ref{eq:17}, \ref{eq:28}, and \ref{eq:29} give \begin{equation} R=-3X_rH_{r\phi r}/X+6\partial_rH_{\phi r\phi}/X=-X_{rr}/X+X_r^2/2X^2, \label{eq:30} \end{equation}} %\indent integrating \begin{equation} H_{\phi r\phi}=-X_r/6+l\sqrt{X}/6, \label{eq:31} \end{equation}} %\indent where $l$ is a constant. From ref{eq:27}, \ref{eq:28}, and \ref{eq:31} and inserting the gauge vector we have \begin{eqnarray} &&H=\frac{k\sqrt{X}}{2}+\frac{l}{12\sqrt{X}}-\frac{X_r}{12X}+\chi_r,\nonumber\\ &&H=\frac{k\sqrt{X}}{2}-\frac{l}{12\sqrt{X}}+\frac{X_r}{12X}-\chi_r,\nonumber\\ &&H=\frac{l\sqrt{X}}{6}-\frac{X_r}{6}-X\chi_r. \label{eq:32} \end{eqnarray} the result is in the Lanczos algebraic gauge when $\chi_r=0$. The derivative of the metric $X_r$ appears in each term. No matter what the choice of algebraic gauge (i.e. choice of $\chi_r$) we cannot remove it. Thus the continuity or otherwise of the Lanczos tensor depends on the continuity or otherwise of the derivative of the metric. We can make any possible discontinuity in any single component, or even a whole component vanish by means of a suitable algebraic gauge. For example choosing \begin{equation} \chi_r=\frac{-k\sqrt{X}}{2}-\frac{l}{12\sqrt{X}}+\frac{X_r}{12X}, \label{eq:33} \end{equation}} %\indent gives \begin{eqnarray} &&H_{trt}=0,\nonumber\\ &&H_{zrz}=k\sqrt{X},\nonumber\\ &&H_{\phi r\phi}=\frac{kX\sqrt{X}}{2}+\frac{l\sqrt{X}}{4}-\frac{X_r}{4}. \label{eq:34} \end{eqnarray} Notice that the Lanczos tensor does not necessarily vanish for flat space-time. For example in the Lanczos algebraic gauge or in the gauge \ref{eq:33} the choice \begin{equation} k=0,~~~ l=2, \label{eq:35} \end{equation}} %\indent gives flat space-time with vanishing Lanczos tensor; however, for example, in the gauge $\chi_r=-X_r/6X$, the Lanczos tensor cannot be made to vanish for flat space-time. From \ref{eq:32} the metric can be expressed in terms of the Lanczos tensor \begin{equation} X=(H^{~t}_{r.t}-H^{~z}_{r.z})^2/k^2. \label{eq:36} \end{equation}} %\indent The differential gauge \ref{eq:9} is \begin{eqnarray} &&L_{tr}=-\partial_rH_{rtr}+X_r(H_{\phi t\phi}/X -H_{rtr})/2X,\nonumber\\ &&L_{r\phi}=\partial_rH_{t\phi r}+X_rH_{tr\phi}/2X,\nonumber\\ &&L_{tz}=\partial_rH_{tzr}+X_rH_{trz}/2X,\nonumber\\ &&L_{r\phi}=\partial_rH_{r\phi r},\nonumber\\ &&L_{rz}=\partial_rH_{rzr}+X_r(-H_{\phi z\phi}/X-H_{rzr}H)/2X,\nonumber\\ &&L_{\phi z}=\partial_rH_{\phi zr}+X_rH_{rz\phi}/2X. \label{eq:37} \end{eqnarray} For the Lanczos differential gauge condition \ref{eq:9}, \ref{eq:37} can be expressed in the integral form \begin{eqnarray} &&H_{r\phi r}=a_1,\nonumber\\ &&H_{ztr}=\frac{a_2}{\sqrt{X}},\nonumber\\ &&H_{\phi tr}=\frac{a_3}{\sqrt{X}},\nonumber\\ &&H_{rtr}=\frac{1}{\sqrt{X}}(a_4+\int X_rX^{-3/2}H_{\phi t\phi}dr),\nonumber\\ &&H_{rzr}=\frac{1}{\sqrt{X}}(a_5+\int X_rX^{-3/2}H_{\phi z\phi}dr),\nonumber\\ &&H_{z\phi r}=\frac{1}{\sqrt{X}}(a_6+\int X_rX^{-3/2}H_{r\phi z}dr), \label{eq:38} \end{eqnarray} where $a_1\ldots a_6$ are constants. This illustrates a property of the differential gauge alluded to in section II; the components of the Lanczos tensor in \ref{eq:38} do not coincide with any of those in \ref{eq:32}, thus these components do not participate in the construction of the Weyl tensor. Here the Lanczos tensor cannot be expressed as the gradient of an anti-symmetric tensor of the second order. Using that the space-time is only $r$ dependent and that the Christoffel symbols are \ref{eq:14}, \ref{eq:10} gives \begin{equation} F_{tr;t}=F_{zr;z}=F_{\phi t;\phi}=F_{rt;t}=F_{rz;z}=F_{r\phi;\phi}=0. \label{eq:39} \end{equation}} %\indent Any added current to \ref{eq:10} would have component $J_z$, and then the cyclic identity \ref{eq:5} would fail. \section{The Covariant Derivative.} \label{sec:IV} In the Aharonov-Bohm effect, the electro-magnetic field alters the dynamics of test particles because the electro-magnetic covariant derivative replaces the partial derivative in the test particles Schr\"odinger's equation. In this section we list $15$ possible covariant derivatives involving the Lanczos tensor. None of the possibilities can be used to complete our analogy with the Aharonov-Bohm effect. This is because the criteria of continuity cannot be used to fix the algebraic gauge in the example in the last section, and all the possibilities give different results depending on algebraic gauge. We denote the covariant derivative by $D_a$ and the coupling constant by $c$. \begin{equation} i)~~~\partial_a\rightarrow D_a=+cH^{~b}_{a.b}. \label{eq:40} \end{equation}} %\indent From ref{eq:7} and \ref{eq:8} we see immediately that this choice depends on the algebraic gauge. The choice \begin{equation} ii)~~~\partial_a\rightarrow D_a=+cH^b_{.ab}, \label{eq:41} \end{equation}} %\indent is the same as \ref{eq:40} with the sign of $c$ reversed, by virtue of the symmetry \ref{eq:4}. The choice \begin{equation} iii)~~~\partial_a\rightarrow D_a=+cH^b_{~ba}, \label{eq:42} \end{equation}} %\indent gives that the covariant and partial derivative are identical by \ref{eq:4}. \begin{equation} iv)~~~\partial_a\rightarrow D_a=+cH^{~B}_{a.B}, \label{eq:43} \end{equation}} %\indent where $B$ is a fixed component not summed. Covariant derivatives of this type have all the disadvantages of \ref{eq:40} to \ref{eq:42}, with the added disadvantage of picking out one component. \begin{equation} v)~~~\partial_a\rightarrow D_a=\partial_a+c(3H^{~t}_{a.t}-H^{~b}_{a.b}, \label{eq:44} \end{equation}} %\indent for our example this is invariant under the choice of algebraic gauge, however by \ref{eq:8} \begin{equation} H^{~b}_{a.b}=3\chi_r, \label{eq:45} \end{equation}} %\indent thus this choice amounts to no more than an arbitrary choice of component in the Lanczos algebraic gauge. \begin{equation} vi)~~~\partial_a\rightarrow D_a=\partial_a+c\epsilon_{abcd}H^{bcd}_{...}, \label{eq:46} \end{equation}} %\indent in our example, or more generally in any space-time which can be expressed with a diagonal metric, components of the Lanczos tensor with a identical adjacent indices plays an essential role, but they would not effect this covariant derivative. \begin{equation} vii)~~~\partial_a\rightarrow D_a=\partial_a+cH_{abc}p^{bc}_{..}, \label{eq:47} \end{equation}} %\indent where $p^{bc}_{..}$ is the stress tensor of the "test" particle. This coupling is unusual as the test particles own stress contributes, it is no longer just a test particle. By changing the algebraic gauge the contribution of $p^{bc}_{..}$ changes in an arbitrary manner and thus this choice is un-useable. viii)Require that the covariant derivative coincides with the weak field covariant derivative when the fields are weak. The weak field covariant derivative is \begin{equation} 2\eta_{ad}W^{.d}_{b.c}=\partial_bh_{ac}+\partial_ch_{ab}-\partial_ah_{bc}. \label{eq:48} \end{equation}} %\indent Using the algebraic gauge \begin{equation} \chi_a=-\frac{1}{6}h_a, \label{eq:49} \end{equation}} %\indent \ref{eq:12} becomes \begin{equation} 4H_{abc}=\partial_bh_{ca}-\partial_ah_{bc}, \label{eq:50} \end{equation}} %\indent using \ref{eq:4} this is \begin{equation} \partial_ah_{bc}=-2H_{abc}. \label{eq:51} \end{equation}} %\indent Substituting \ref{eq:50} into \ref{eq:47} and using \ref{eq:4} and \ref{eq:5} gives \begin{equation} 2\eta_{ad}W^{.d}_{b.c}=4H_{abc}+2H_{acb}, \label{eq:52} \end{equation}} %\indent now \begin{equation} \eta_{ad}W^{~d}_{c.b}=\eta_{ad}W^{~d}_{b.c}, \label{eq:53} \end{equation}} %\indent thus from \ref{eq:51} and \ref{eq:52} \begin{equation} H_{abc}=H_{acb}, \label{eq:54} \end{equation}} %\indent using \ref{eq:4} and \ref{eq:5}, \ref{eq:54} gives \begin{equation} H_{bca}=0. \label{eq:55} \end{equation}} %\indent Thus the weak field covariant derivative cannot be expressed in terms of the weak field expression for $H_{abc}$, and we cannot require that the covariant derivative coincides with the weak field covariant derivative when the fields are weak. ix)Apply $H_{abc}=F_{ab;c}$ and then use the electro-magnetic covariant derivative. This is too restrictive, there is no $F_{ab}$ satisfying this in our example. \begin{equation} x)~~~\partial_a\rightarrow D_a=\partial_a+cH_a, \label{eq:56} \end{equation}} %\indent where \begin{equation} H_a=H_{abc}H^{bc}_{..}, \label{eq:57} \end{equation}} %\indent and $H^{bc}_{..}$ is given by \ref{eq:6}. This has the disadvantage that it involves products and derivatives of the Lanczos tensor; the electro-magnetic covariant derivative has no products and derivatives of the vector potential. For the example of the previous section in the Lanczos algebraic gauge \begin{equation} H_r=-\frac{k^2X_r}{2}-\frac{X_rX_{rr}}{24X^2}+\frac{5X_r^3}{144X^3} +\frac{l^2X_r}{72X^2}-\frac{7lX_r^2}{144X^{3/2}}+\frac{lX_{rr}}{24X^{3/2}}, \label{eq:58} \end{equation}} %\indent in the $\chi_r=X_r/12X$ gauge \begin{equation} H_r=-\frac{k^2X_r}{2}-\frac{X_rX_{rr}}{16X^2}+\frac{X_r^3}{32X^3} +\frac{l^2X_r}{72X^2}-\frac{lX^2_r}{24X^{3/2}}+\frac{lX_{rr}}{24X^{3/2}}, \label{eq:59} \end{equation}} %\indent in the $\chi_r=-X_r/6X$ gauge \begin{equation} H_r=-\frac{k^2X_r}{2}-\frac{X_rX_{rr}}{8X^2}+\frac{X_r^3}{16X^3} +\frac{l^2X_r}{72X^2}-\frac{lX_r^2}{X^{3/2}}+\frac{lX_{rr}}{24X^{3/2}}, \label{eq:60} \end{equation}} %\indent illustrating the algebraic gauge dependence of $H_a$. xi)Replacing ref{eq:57} by \begin{equation} H_a=H_{bac}H^{bc}_{..}, \label{eq:61} \end{equation}} %\indent is the same as \ref{eq:57} with the sign of the coupling constant reversed, xii)replacing \ref{eq:57} by \begin{equation} H_a=H_{bac}H^{bc}_{..}, \label{eq:62} \end{equation}} %\indent we have that the indices $b$ and $c$ will always be identical for any space-time where the metric is in diagonal form and thus $H_a=0$. Require that rather than a covariant derivative we need a change of phase \begin{equation} xiii)~~~\alpha\rightarrow\alpha'=\alpha+cH_{abc}H^{abc}_{...}, \label{eq:63} \end{equation}} %\indent This has the same difficulties as the example of the preceeding paragraph, and also we have a difficulty in how to integrate over the different trajectories. xiv)Another possibility is to start with the Klein-Gordon equation or the Dirac equation in the space-time \ref{eq:13}. In the non-relativistic limit the Schr\"odinger equation can be recovered from the Klein-Gordon and Dirac equations and this might give information on the correct covariant derivative. To investigate this we begin by showing that there are no solutions with non-vanishing gravitational and Klein-Gordon field with line element \ref{eq:13}. The equations for an Einstein-Klein-Gordon field are \begin{equation} R_{ab}=2\phi_a\phi_b+2g_{ab}m^2\phi^2, \label{eq:64} \end{equation}} %\indent and \begin{equation} m^2\phi=\Box\phi=\frac{1}{\sqrt{-g}}(\sqrt{-g}g^{ab}_{..}\phi_a)_b. \label{eq:65} \end{equation}} %\indent Thus for the line element \ref{eq:13}, using $ph_\phi=0$, \begin{equation} R^r_{.r}=2\phi'^2+2m^2\phi^2,~~~ R^{\phi}_{.\phi}=2m^2\phi^2, \label{eq:66} \end{equation}} %\indent Now \ref{eq:15} gives $R^r_{.r}=R^\phi_{.\phi}$ therefore $\phi'=0$, integrating \begin{equation} \phi=\sigma, \label{eq:67} \end{equation}} %\indent where $\sigma$ is a constant. Equation \ref{eq:65} becomes \begin{equation} m^2\phi=\frac{1}{\sqrt{X}}(\sqrt{X}\phi')', \label{eq:68} \end{equation}} %\indent which vanishes by \ref{eq:67}. Therefore $m^2=0$ and $\phi=\sigma$ or $\phi=0$. Thus there are no solutions of \ref{eq:64} with line element \ref{eq:13} and both $m$ and $\phi$ non-vanishing. Similarly, using \cite{bi:PR} it can be shown that there are no solutions of the Einstein-Dirac equations with line element \ref{eq:13} and both gravitational and Dirac fields non-vanishing. Instead of considering coupled systems we could consider the Klein-Gordon or Dirac fields as test fields which do not contribute to the stress of the space-time. Using \ref{eq:37}, \ref{eq:65}, and \ref{eq:68} the Klein-Gordon equation can be expressed as \begin{equation} m^2\phi=\nabla_a\partial^a_.=\phi"+(ln(H^{~t}_{r.r}-H^{~z}_{r.z}))'\phi', \label{eq:69} \end{equation}} %\indent suggesting the covariant derivative \begin{equation} xv)~~~\partial_a\rightarrow D_a=\partial_a+(ln(H^{~t}_{r.t}-H^{~z}_{r.z}))'. \label{eq:70} \end{equation}} %\indent This covariant derivative has the disadvantage of artificially picking out a component and involves neither the gauge vector $\chi_a$ or the constants $k$ and $l$. The analogy with the Aharonov-Bohm effect suggests that the exterior region should be Minkowski space-time, where $X=r^2$, and that the criteria of continuity should fix the gauge vector $\chi_a$ and the constants $k$ and $l$; but this cannot be done with covariant derivative \ref{eq:70}. \section{Conclusion} \label{sec:V} An attempt was made to test in principle whether the Lanczos tensor is microscopically dynamically significant in the quantum realm in a similar manner to the vector potential. So far the results have not produced a definitive result: however whether it is possible to quantize by this method should become clearer upon the discovery of a suitable exact solution to the general relativity field equations which can produce a closer analogy to the Aharonov-Bohm experiment.
\section{Introduction} A correct description of both the ground and excited states of scalar mesons encounters a variety of complex problems. Let us point out some of them. i) For a long time, the experimental status of the lightest scalar isoscalar singlet meson was unclear. In some papers, the resonance $f_0(1370)$ was considered as that particle \cite{F1370}, and it was not until 1998 that the resonance $f_0(400-1200)$ was included into the summary tables of PDG review% \footnote{ However, in earlier editions of PDG the light $\sigma$ state still could be found; it was excluded later. } \cite{PDG}. ii) The scalar isoscalar states, as having quantum numbers of vacuum, are most probably get mixed with glueballs \cite{glueball}. iii) There is also a lot of problems related to the description of $f_0(980)$ and $a_0(980)$. Their unusual experimental branching ratios for several decays have brought forth different ideas concerning the structure of the mesons. Among them, there are the quark-antiquark model \cite{F1370,glueball,volk}, the four-quark model \cite{achasov} and the kaon molecule model \cite{kmolecule}. iv) The strange meson $K^*_0(1430)$ seems too heavy to be the ground state: $1{\kern2mm\rm GeV}$ is more characteristic of the ground meson states (see \cite{ishida}). The description of the ground and excited states of the pion, kaon and the vector meson nonet in the framework of a nonlocal version of the NJL model has been done in our earlier papers \cite{weiss,volk97,ven}. Here we intend to study the ground and first radially excited states of $\eta$, $\eta'$ mesons and of the scalar meson nonet. To produce correct masses for the ground states of $\eta$ and $\eta'$, we, as usual, introduce 't~Hooft interaction \cite{klev,volk98}. Although our model is nonlocal, which is reflected in the presence of form factors in the four-quark vertices, we, nevertheless, assume the 't~Hooft term local. The form factors in scalar channels of quark current-current interaction are chosen identical to those in the pseudoscalar channel. This is a requirement of the global chiral symmetry of quark interaction. With that assumption, there is no need for additional parameters in the form factors of scalar quark vertices. Therefore, the masses of scalar mesons can be immediately predicted after fixing the form factor parameters by the pseudoscalar meson masses from experiment. As a result, we have found that the model masses of the radial excitations of scalar isoscalar mesons are close to the experimentally observed $f_0(1370)$, $f_J(1710)$% \footnote{ We assume hereafter $f_J(1710)$ is an isoscalar ($J=0$). }% , $a_0(1450)$ mesons. This allows us to interpret them as the first radial excitations of mesons $f_0(400-1200)$, $f_0(980)$ and $a_0(980)$. As to the state $f_0(1500)$, we are inclined to consider it as a glueball. In our further works we will take into account possible mixing of the $\bar qq$ scalar meson states with glueballs \cite{glueball}. Concerning the strange scalar $K_{0}^{*}$, we think that the state with mass $1430{\kern2mm\rm MeV}$ is much likely a radial excitation of a light and wide resonance with mass about $960{\kern2mm\rm MeV}$ (see \cite{ishida}). Further discussion on this problem is given in conclusion. As to the ground states $a_0(980)$ and $f_0(980)$, the detailed discussion on their internal structure and properties is beyond the scope of our paper. Our paper is organized as follows. In Sec.~2, we introduce the chiral quark Lagrangian with nonlocal four-quark vertices and local 't Hooft interaction. In Sec.~3, we calculate the effective Lagrangian for isovector and strange mesons in the one-loop approximation. There we renormalize meson fields and transform the free part of the Lagrangian to the diagonal form and obtain meson mass formulae. Section 4 is devoted to isoscalar mesons where we find masses and mixing coefficients. The model parameters are discussed in Sec.~5. In Sec.~6, we calculate the widths of major strong decays of excited states of $a_0$, $\sigma$ and $f_0$ mesons. The results of our work and possible ways to improve the model are discussed in Sec.~7. Some details of the calculations fulfilled in Sec.~4 and 6 are given in Appendices A and B. \section{$U(3)\times U(3)$ chiral Lagrangian with excited meson states and 't Hooft interaction} We use a nonlocal separable four-quark interaction of a current-current form which admits nonlocal vertices (form factors) in the quark currents, and a pure local six-quark 't Hooft interaction \cite{klev,volk98}: \begin{eqnarray} {\cal L}(\bar q, q) &=& \int\! d^4x\; \bar q(x) (i \partial\hspace{-.5em}/\hspace{.15em} -m^0) q(x)+ {\cal L}^{(4)}_{\rm int}+ {\cal L}^{(6)}_{\rm int}, \label{lag}\\ {\cal L}^{(4)}_{\rm int} &=& \int\! d^4x\sum^{8}_{a=0}\sum^{N}_{i=1} \frac{G}{2}[j_{S,i}^a(x) j_{S,i}^a(x)+ j_{P,i}^a(x) j_{P,i}^a(x)],\\ {\cal L}^{(6)}_{\rm int}&=&-K \left[\det \left[\bar q (1+\gamma_5)q\right]+ \det\left[\bar q (1-\gamma_5)q\right] \right]. \end{eqnarray} Here, $m^0$ is the current quark mass matrix ($m_u^0\approx m_d^0$) and $j^a_{S(P),i}$ denotes the scalar (pseudoscalar) quark currents \begin{equation} j^a_{S(P),i}(x)= \int\! d^4x_1 d^4x_2\; \bar q(x_1) F^a_{S(P),i }(x;x_1,x_2) q(x_2) \end{equation} where $ F^a_{S(P),i}(x;x_1,x_2)$ are the scalar (pseudoscalar) nonlocal quark vertices. To describe the first radial excitations of mesons, we take the form factors in momentum space as follows (see \cite{weiss,volk97,ven}), \begin{equation} \begin{array}{ll} F_{S,j}^a({\bf k})=\lambda^a f^a_j,\quad & F_{P,j}^a= i\gamma_5 \lambda^a f^a_j \end{array} \end{equation} \begin{equation} f^a_1\equiv 1,\quad f^a_2\equiv f_a({\bf k})=c_a(1+d_a {\bf k}^2),\label{fDef} \end{equation} where $\lambda^a$ are Gell--Mann matrices, $\lambda^0 = {\sqrt{2\over 3}}${\bf 1}, with {\bf 1} being the unit matrix. Here, we consider the form factors in the rest frame of mesons \footnote{The form factors depend on the transversal parts of the relative momentum of quark-antiquark pairs $k_{\perp} = k - \frac{k\cdot P}{P^2}P$, where $k$ and $P$ are the relative and total momenta of a quark-antiquark pair, respectively. Then, in the rest frame of mesons, ${\bf P}_{meson}$ = 0, the transversal momentum is $k_{\perp } = (0, {\vec k})$, and we can define the form factors as depending on the 3-dimensional momentum ${\vec k}$ alone. }% . The part of the Lagrangian (\ref{lag}), describing the ground states and first radial excitations, can be rewritten in the following form (see \cite{klev} and \cite{volk98}): \begin{eqnarray} {\cal L}&=& \int\! d^4x \biggl\{ \bar q(x) (i\partial\hspace{-.5em}/\hspace{.15em}-m^0) q(x) + \frac{G}{2}\sum_{a=0}^8 \left[\left(j_{S,2}^a\right)^2+ \left(j_{P,2}^a\right)^2\right]+ \nonumber\\ && \frac12\sum_{a=1}^9\left[G^{(-)}_a \left(\bar q(x) \tau_a q(x)\right)^2+ G^{(+)}_a\left(\bar q(x)i\gamma_5 \tau_a q(x)\right)^2\right]+ \label{lagr}\\ && G_{us}^{(-)}(\bar q(x)\lambda_u q(x)) (\bar q(x)\lambda_s q(x)) + G_{us}^{(+)}(\bar q(x)i\gamma_5\lambda_u q(x)) (\bar q(x)i\gamma_5\lambda_s q(x))\biggr\},\nonumber \end{eqnarray} where \begin{eqnarray} &&{\tau}_i={\lambda}_i ~~~ (i=1,...,7),~~~ \tau_8 = \lambda_u = ({\sqrt 2} \lambda_0 + \lambda_8)/{\sqrt 3},\nonumber\\ && \tau_9 = \lambda_s = (-\lambda_0 + {\sqrt 2}\lambda_8)/{\sqrt 3}, \label{DefG} \\ && G_1^{(\pm)}=G_2^{(\pm)}=G_3^{(\pm)}= G \pm 4Km_sI_1(m_s), \nonumber \\ && G_4^{(\pm)}=G_5^{(\pm)}=G_6^{(\pm)}= G_7^{(\pm)}= G \pm 4Km_uI_1(m_u), \nonumber \\ && G_u^{(\pm)}= G \mp 4Km_sI_1(m_s), ~~~ G_s^{(\pm)}= G, ~~~ G_{us}^{(\pm)}= \pm 4{\sqrt 2}Km_uI_1(m_u).\nonumber \end{eqnarray} Here $m_u$ and $m_s$ are the constituent quark masses and $I_1(m_q)$ is the integral which for an arbitrary $n$ is defined as follows \begin{equation} I_n(m_q)={-i N_c\over (2\pi)^4} \int_{\Lambda_3}\!d^4 k {1\over (m^2_q-k^2)^n} . \label{DefI} \end{equation} The 3-dimensional cut-off $\Lambda_3$ in (\ref{DefI}) is implemented to regularize the divergent integrals% \footnote{For instance, $I_1(m)=\frac{N_c m^2}{8\pi^2}\left.[x\sqrt{x^2+1}-\ln(x+\sqrt{x^2+1})]\right|_{x=\Lambda_3/m}$. }% . \section{The masses of isovector and strange mesons (ground and excited states)} After bosonization, the part of Lagrangian (\ref{lagr}), describing the isovector and strange mesons, takes the form \begin{eqnarray} && {\cal L}(a_{0,1},K_0^*{}_{,1},\pi_1,K_1, a_{0,2}, K_0^*{}_{,2}, \pi_2, K_2)= -\frac{a_{0,1}^2}{2G_{a_0}}-\frac{{K_0^*{}_{,1}}^2}{G_{K_0^*}}-\frac{\pi_1^2}{2G_\pi}- \frac{K_1^2}{G_K}-\nonumber\\ && \frac{1}{2G} ( a_{0,2}^2+2(K_0^*{}_{,2})^2+ \pi_2^2+2 K_2^2)-\nonumber\\ && i N_c {\rm Tr}\ln\left[1+ \frac{1}{i\partial\hspace{-.5em}/\hspace{.15em}-m}\sum_{a=1}^7\sum_{j=1}^2 \lambda_a\left[ \sigma^a_j+i\gamma_5\varphi^a_j\right]f^a_j \right] \label{bosLag} \end{eqnarray} where $m={\rm diag}(m_u,m_d,m_s)$ is the matrix of constituent quark masses ($m_u\approx m_d$), $\sigma^a_j$ and $\phi^a_j$ are the scalar and pseudoscalar fields: $\sum_{a=1}^{3}(\sigma^a_j)^2\equiv a_{0,j}^2=({a_{0,j}^0})^2+2a_{0,j}^+a_{0,j}^-$, $\sum_{a=4}^{7}(\sigma^a_j)^2\equiv 2{K_{0}^{*}{}_{,j}}^2= 2(\bar{K_{0}^{*}}_{,j})^0(K_{0}^{*}{}_{,j})^0+2(K_{0}^{*}{}_{,j})^+(K_{0}^{*}{}_{,j})^-$, $\sum_{a=1}^{3}(\varphi^a_j)^2\equiv \pi_j^2=({\pi_j^0})^2+2\pi_j^+\pi_j^-$, $\sum_{a=4}^{7}(\varphi^a_j)^2\equiv 2K_j^2=2\bar K^0_j K^0_j+2\bar K^+_jK^-_j$. As to the coupling constants $G_a$, they will be defined later (see Sect.~5 and (\ref{DefG})). The free part of Lagrangian (\ref{bosLag}) has the following form \begin{eqnarray} {\cal L}^{(2)}(\sigma,\varphi)=\frac12\sum_{i,j=1}^2\sum_{a=1}^7 \left(\sigma^a_i K_{\sigma,ij}^a(P)\sigma^a_j+ \varphi^a_i K_{\varphi,ij}^a(P)\varphi^a_j\right) \label{L2} \end{eqnarray} where the coefficients $K^{a}_{\sigma(\varphi), ij}(P)$ are given below, \begin{eqnarray} && K_{\sigma(\varphi), ij}^{a}(P)= -\delta_{ij}\left[\frac{\delta_{i1}}{G_a^{(\mp)}}+\frac{\delta_{i2}}{G}\right]-\nonumber\\ && i N_c {\rm Tr} \int_{\Lambda_3}\frac{d^4k}{(2\pi)^4} {1\over k\hspace{-.5em}/\hspace{.15em}+P\hspace{-.5em}/\hspace{.15em}/2-m^a_q} r^{\sigma(\varphi)}f_i^a {1\over k\hspace{-.5em}/\hspace{.15em}-P\hspace{-.5em}/\hspace{.15em}/2-m^a_{q'}} r^{\sigma(\varphi)}f_j^a, \label{K_full} \end{eqnarray} \begin{equation} r^\sigma=1,\quad r^{\phi}=i\gamma_5, \end{equation} \begin{equation} m_q^a = m_u~~(a = 1,...,7);\quad m_{q'}^a = m_u~~(a = 1,...,3);~~ m_{q'}^a = m_s~~ (a = 4,...,7), \label{m_q^a} \end{equation} with $m_u$ and $m_s$ being the constituent quark masses and $f_j^a$ defined in (\ref{fDef}). Integral (\ref{K_full}) is evaluated by expanding in the meson field momentum $P$. To order $P^2$, one obtains \begin{eqnarray} K_{\sigma(\varphi),11}^a(P)&=& Z_{\sigma(\varphi),1}^a (P^2 - (m_q^a\pm m_{q'}^a)^2- M_{\sigma^a(\varphi^a),1}^2 ),\nonumber\\ K_{\sigma(\varphi),22}^a(P) &=& Z_{\sigma(\varphi), 2}^a (P^2 - (m_q^a\pm m_{q'}^a)^2- M_{\sigma^a(\varphi^a),2}^2 ), \nonumber \\ K_{\sigma(\varphi),12}^a(P) &=& K_{\sigma(\varphi),21}^a(P) \;\; = \;\; \gamma_{\sigma(\varphi)}^a (P^2 - (m^a_q\pm m^a_{q'})^2 ), \label{Ks_matrix} \end{eqnarray} where \begin{equation} Z_{\sigma,1}^a = 4 I_2^a, \hspace{2em} Z_{\sigma,2}^a = 4 I_2^{ff a}, \hspace{2em} \gamma_{\sigma}^a = 4 I_2^{f a}, \label{Zs} \end{equation} \begin{equation} Z_{\varphi,1}^a = Z Z_{\sigma,1}^a, \hspace{2em} Z_{\varphi,2}^a = Z_{\sigma,2}^a , \hspace{2em} \gamma_{\varphi}^a = Z^{1/2}\gamma_{\sigma}^a \label{Zp} \end{equation} and \begin{eqnarray} M_{\sigma^a(\varphi^a),1}^2 &=& (Z_{\sigma(\varphi),1}^a)^{-1} \left[\frac{1}{G_a^{(\mp)}}-4(I_1(m_q^a) + I_1(m_{q'}^a))\right] \label{M_1} \\ M_{\sigma^a(\varphi^a),2}^2 &=& (Z_{\sigma(\varphi),2}^a)^{-1} \left[\frac{1}{G}-4(I_1^{ff a}(m_q^a) + I_1^{ff a}(m_{q'}^a))\right]. \label{M_2} \end{eqnarray} The factor $Z$ here appears due to account of $\pi-a_1$-transitions~\cite{volk,volk97}, \begin{equation} Z=1-\frac{6 m_u^2}{ M_{a_1}^2}, \end{equation} and the integrals $I_2^{f..f}$ contain form factors: \begin{equation} I_2^{f..f_a}(m^a_q,m^a_{q'})={-i N_c\over (2\pi)^4} \int_{\Lambda_3} d^4 k {f_a({\bf k})..f_a({\bf k})\over ((m^a_q)^2-k^2)((m^a_{q'})^2-k^2)}. \label{DefIf} \end{equation} Further, we consider only the scalar isovector and strange mesons because the masses of the pseudoscalar mesons have been already described in \cite{volk97}. After the renormalization of the scalar fields \begin{equation} \sigma_i^{a r}=\sqrt{Z_{\sigma,i}^a} \sigma_i^{a} \label{renorm} \end{equation} the part of Lagrangian (\ref{L2}) which describes the scalar mesons takes the form \begin{eqnarray} {\cal L}^{(2)}_{a_0}&=&\frac12 \left( P^2-4 m^2_u -M^2_{a_0, 1}\right)a_{0, 1}^2+ \Gamma_{a_0}\left( P^2-4m_u^2\right)a_{0, 1}a_{0, 2}+\nonumber\\ &&\frac12\left(P^2-4m_u^2- M_{a_{0},2}^2\right)a_{0, 2}^2, \label{La0} \end{eqnarray} \begin{eqnarray} {\cal L}^{(2)}_{K_{0}^{*}}\!&=&\!\frac12\! \left( P^2-(m_u+ m_s)^2\! -\! M^2_{K_{0}^{*},1}\right)K_{0}^{*}{}_{,1}^2\! +\! \Gamma_{K_{0}^{*}} \left(P^2-(m_u+m_s)^2\right)K_{0}^{*}{}_{,1}K_{0}^{*}{}_{,2}+\nonumber\\ &&\!\frac12\left(P^2-(m_u+m_s)^2- M_{K_{0}^{*}{},2}^2\right)K_{0}^{*}{}_{,2}^2, \label{LK} \end{eqnarray} where \begin{equation} \Gamma_{\sigma^a}=\frac{I_2^{f_a}}{\sqrt{I_2 I_2^{ff_a}}}.\qquad \end{equation} After the transformations of the meson fields \begin{eqnarray} \sigma^a &=& \cos( \theta_{\sigma,a} - \theta_{\sigma,a}^0) \sigma_1^{ar} - \cos( \theta_{\sigma,a} + \theta_{\sigma,a}^0) \sigma_2^{ar}, \nonumber \\ \hat\sigma^{a} &=& \sin ( \theta_{\sigma,a} - \theta_{\sigma,a}^0) \sigma_1^{ar} - \sin ( \theta_{\sigma,a} + \theta_{\sigma,a}^0) \sigma_2^{ar}, \label{transf} \end{eqnarray} Lagrangians (\ref{La0}) and (\ref{LK}) take the diagonal form: \begin{eqnarray} L_{a_0}^{(2)} &=& {\textstyle{\frac{1}{2}}} (P^2 - M_{a_0}^2)~ a_0^2 + {\textstyle{\frac{1}{2}}} (P^2 - M_{\hat a_0}^2)\hat a_0^{ 2}, \\ L_{K_{0}^{*}}^{(2)} &=& {\textstyle{\frac{1}{2}}} (P^2 - M_{K_{0}^{*}}^2)~ K_{0}^{*}{}^2 + {\textstyle{\frac{1}{2}}} (P^2 - M_{\hatK_{0}^{*}}^2)\hat K_{0}^{*}{}^{ 2}. \label{L_pK} \end{eqnarray} Here we have \begin{eqnarray} && M^2_{(a_0, \hat a_0)} = \frac{1}{2 (1 - \Gamma^2_{a_0})} \biggl[M^2_{a_{0}, 1} + M^2_{a_{0}, 2}\pm \nonumber \\ && \qquad \sqrt{(M^2_{a_{0}, 1} - M^2_{a_{0}, 2})^2 + (2 M_{a_{0}, 1} M_{a_{0}, 2} \Gamma_{a_0})^2}\biggr]+4m_u^2, \\ && M^2_{(K_{0}^{*}, \hatK_{0}^{*})} = \frac{1}{2 (1 - \Gamma^2_{K_{0}^{*}})} \biggl[M^2_{K_{0}^{*}, 1} + M^2_{K_{0}^{*}, 2}\pm \nonumber \\ && \qquad \sqrt{(M^2_{K_{0}^{*},1} - M^2_{K_{0}^{*},2})^2 + (2 M_{K_{0}^{*},1} M_{K_{0}^{*},2} \Gamma_{K_{0}^{*}})^2}\biggr]+ (m_u+m_s)^2, \label{MpKstar} \end{eqnarray} and \begin{equation} \tan 2 {\bar{\theta}}_{\sigma,a} = \sqrt{\frac{1}{\Gamma_{\sigma^a}^2} - 1}~\left[ \frac{M_{\sigma^a,1}^2- M_{\sigma^a,2}^2}{M_{\sigma^a,1}^2 + M_{\sigma^a,2}^2} \right],\qquad 2 \theta_{\sigma,a} = 2 {\bar{\theta}}_{\sigma,a} + \pi, \label{tan} \end{equation} \begin{equation} \sin \theta_{\sigma,a}^{0} =\sqrt{{1+\Gamma_{\sigma^a}}\over 2}. \label{theta0} \end{equation} The caret symbol stands for the first radial excitations of mesons. Transformations (\ref{transf}) express the ``physical'' fields $\sigma$ and $\hat\sigma$ through the ``bare'' ones $\sigma^{ar}_i$ and for calculations, these equations must be inverted. For practical use, we collect the values of the inverted equations for the scalar and pseudoscalar fields% \footnote{ Although the formulae for the pseudoscalars are not displayed here (they have been already obtained in \cite{volk97}) we need the values because we are going to calculate the decay widths of processes where pions and kaons are secondary particles. } in Table \ref{mixingTable}. \begin{table} \caption{The mixing coefficients for the ground and first radially excited states of the scalar and pseudoscalar isovector and strange mesons. The caret symbol marks the excited states.} \label{mixingTable} $$ \begin{array}{||r|c|c||} \hline & a_0 & \hat a_0\\ \hline a_{0,1} &0.87 &0.82 \\ a_{0,2} &0.22 &-1.17 \\ \hline \end{array} \quad \begin{array}{||r|c|c||} \hline & K_{0}^{*} & \hatK_{0}^{*}\\ \hline K_{0}^{*}{}_{,1} &0.83 &0.89 \\ K_{0}^{*}{}_{,2} &0.28 &-1.11 \\ \hline \end{array} $$ $$ \begin{array}{||r|c|c||} \hline & \pi & \hat\pi\\ \hline \pi_1 &1.00 &0.54 \\ \pi_2 &0.01 &-1.14 \\ \hline \end{array} \quad \begin{array}{||r|c|c||} \hline & K & \hat K \\ \hline K_1 &0.96 &0.56 \\ K_2 & 0.09 &-1.11 \\ \hline \end{array} $$ \end{table} \section{The masses of isoscalar mesons (the ground and excited states)} The 't Hooft interaction effectively gives rise to the additional four-quark vertices in the isoscalar part of Lagrangian (\ref{lagr}): \begin{equation} {\cal L}_{\rm isosc}=\sum_{a,b=8}^9\left[ (\bar q \tau_a q)T^S_{a b} (\bar q \tau_b q)+ (\bar q i\gamma_5\tau_a q)T^P_{a b} (\bar q i\gamma_5\tau_b q)\right] \end{equation} where $T^{S(P)}$ is a matrix with elements defined as follows (for the definition of $G_u^{(\mp)}$, $G_s^{(\mp)}$ and $G_{us}^{(\mp)}$ see (\ref{DefG})) \begin{equation} \begin{array}{ll} T^{S(P)}_{88}=G^{(\mp)}_{u}/2,\quad & T^{S(P)}_{89}=G^{(\mp)}_{us}/2, \\ T^{S(P)}_{98}=G^{(\mp)}_{us}/2,\quad & T^{S(P)}_{99}=G^{(\mp)}_{s}/2. \end{array} \end{equation} This leads to nondiagonal terms in the free part of the effective Lagrangian for isoscalar scalar and pseudoscalar mesons after bosonization \begin{eqnarray} &&{\cal L}_{\rm isosc}(\sigma,\varphi)= -{1\over 4}\sum_{a,b=8}^9\left[ \sigma^{a}_1 (T^S)^{-1}_{a b}\sigma^{b}_1 + \varphi^{a}_1 (T^P)^{-1}_{a b}\varphi^{b}_1\right]- \nonumber \\ && {1\over 2G}\sum_{a=8}^9 \left[\left(\sigma^{a}_2\right)^2 + \left(\varphi^{a}_2\right)^2 \right]- \nonumber \\ &&i~{\rm Tr}\ln \left\{1 + {1\over i{ \partial\hspace{-.5em}/\hspace{.15em}} - m} \sum_{a=8}^9\sum_{j=1}^2 \tau^{a}[ \sigma^{a}_j + i\gamma_5 \varphi^{a}_j ]f^{a}_j \right\}, \label{Lbar} \end{eqnarray} where $(T^{S(P)})^{-1}$ is the inverse of $T^{S(P)}$: \begin{equation} \begin{array}{ll} (T^{S(P)})^{-1}_{88}=2G^{(\mp)}_{s}/D^{(\mp)},\quad & (T^{S(P)})^{-1}_{89}= (T^{S(P)})^{-1}_{98}=-2G^{(\mp)}_{us}/D^{(\mp)}, \\ (T^{S(P)})^{-1}_{99}=2G^{(\mp)}_{u}/D^{(\mp)},\quad & D^{(\mp)}=G^{(\mp)}_u G^{(\mp)}_s-(G^{(\mp)}_{us})^2 . \end{array} \label{Tps1} \end{equation} From (\ref{Lbar}), in the one-loop approximation, one obtains the free part of the effective Lagrangian \begin{eqnarray} {\cal L}^{(2)}(\sigma,\phi)=\frac12\sum_{i,j=1}^2\sum_{a,b=8}^9 \left(\sigma^a_i K_{\sigma,ij}^{[a,b]}(P)\sigma^b_j+ \varphi^a_i K_{\phi,ij}^{[a,b]}(P)\varphi^b_j\right). \label{Lisosc} \end{eqnarray} The definition of $K_{\sigma(\varphi),i}^{[a,b]}$ is given in Appendix A. After the renormalization of both the scalar and pseudoscalar fields, analogous to (\ref{renorm}), we come to the Lagrangian which can be represented in a form slightly different from that of (\ref{Lisosc}). It is convenient to introduce 4-vectors of ``bare'' fields \begin{equation} \Phi=(\varphi_{1}^{8\,r},\varphi_{2}^{8\,r}, \varphi_{1}^{9\,r},\varphi_{2}^{9\,r}), \qquad \Sigma=(\sigma_{1}^{8\,r},\sigma_{2}^{8\,r}, \sigma_{1}^{9\,r},\sigma_{2}^{9\,r}). \end{equation} Thus, we have \begin{eqnarray} {\cal L}^{(2)}(\Sigma,\Phi)=\frac12\sum_{i,j=1}^4 \left(\Sigma_i {\cal K}_{\Sigma,ij}(P)\Sigma_j+ \Phi_i {\cal K}_{\Phi,ij}(P)\Phi_j\right) \label{L2a} \end{eqnarray} where we introduced new functions ${\cal K}_{\Sigma(\Phi),ij}(P)$ (see Appendix A). Up to this moment one has four pseudoscalar and four scalar meson states which are the octet and nonet singlets. The mesons of the same parity have the same quantum numbers and, therefore, they are expected to be mixed. In our model the mixing is represented by $4\times 4$ matrices $R^{\sigma(\varphi)}$ which transform the ``bare'' fields $\varphi_{i}^{8\,r}$, $\varphi_{i}^{9\,r}$, $\sigma_{i}^{8\,r}$ and $\sigma_{i}^{9\,r}$ entering the 4-vectors $\Phi$ and $\Sigma$ to the ``physical'' ones $\eta$, $\eta'$, $\hat\eta$, $\hat\eta'$, $\sigma$, $\hat\sigma$, $f_0$ and $\hat f_0$ represented as components of vectors $\Phi_{\rm ph}$ and $\Sigma_{\rm ph}$: \begin{equation} \Phi_{\rm ph}=(\eta,\hat\eta,\eta',\hat\eta'), \qquad \Sigma_{\rm ph}=(\sigma,\hat\sigma,f_0,\hat f_0) \end{equation} where, let us remind once more, a caret over a meson field stands for the first radial excitation of the meson. The transformation $R^{\sigma(\varphi)}$ is linear and nonorthogonal: \begin{equation} \Phi_{\rm ph}=R^{\varphi}\Phi,\qquad \Sigma_{\rm ph}=R^{\sigma}\Sigma. \end{equation} In terms of ``physical'' fields the free part of the effective Lagrangian is of the conventional form and the coefficients of matrices $R^{\sigma(\varphi)}$ give the mixing of the $\bar uu$ and $\bar ss$ components, with and without form factors. Because of the complexity of the procedure of diagonalization for the matrices of dimensions greater than 2, there is no such simple formulae as, {\it e.g.}, in (\ref{transf}). Hence, we do not implement it analytically but use numerical methods to obtain matrix elements (see Table~\ref{isoscMixTab}). \begin{table} \caption{The mixing coefficients for the isoscalar meson states} \label{isoscMixTab} $$ \begin{array}{||r|c|c|c|c||} \hline\hline &\eta &\hat\eta &\eta' &\hat\eta'\\ \hline \varphi^8_{1} &0.71 &0.62 &-0.32 &0.56 \\ \varphi^8_{2} &0.11 &-0.87 &-0.48 &-0.54 \\ \varphi^9_{1} &0.62 &0.19 &0.56 &-0.67 \\ \varphi^9_{2} &0.06 &-0.66 &0.30 &0.82 \\ \hline \end{array} $$ $$ \begin{array}{||r|c|c|c|c||} \hline\hline &\sigma &\hat\sigma &f_0 &\hat f_0\\ \hline \sigma^8_{1} &-0.98 &-0.66 &0.10 &0.17 \\ \sigma^8_{2}&0.02 &1.15 &0.26 &-0.17 \\ \sigma^9_{1} &0.27 &-0.09 &0.82 &0.71 \\ \sigma^9_{2}&-0.03 &-0.21 &0.22 &-1.08 \\ \hline \end{array} $$ \end{table} \section{Model parameters and meson masses} In our model we have five basic parameters: the masses of the constituent $u(d)$ and $s$ quarks, $m_u=m_d$ and $m_s$, the cut-off parameter $\Lambda_3$, the four-quark coupling constant $G$ and the 't Hooft coupling constant $K$. We have fixed these parameters with the help of input parameters: the pion decay constant $F_\pi=93 {\kern2mm\rm MeV}$, the $\rho$-meson decay constant $g_\rho=6.14$ (decay $\rho\to2\pi$)% \footnote{Here we do not consider vector and axial-vector mesons, however, we have used the relation $g_\rho=\sqrt{6}g_{\sigma}$ together with the Goldberger--Treiman relation $g_\pi=\frac{m}{F_\pi}=Z^{-1/2}g_{\sigma}$ to fix the parameters $m_u$ and $\Lambda_3$ (see \cite{volk97}). }, the masses of pion and kaon and the mass difference of $\eta$ and $\eta'$ mesons (for details of these calculations, see \cite{volk97,ven,volk98}). Here we give only numerical estimates of these parameters: \begin{eqnarray} &&m_u=280{\kern2mm\rm MeV},\quad m_s=405 {\kern2mm\rm MeV}, \quad\Lambda_3=1.03 {\kern2mm\rm GeV}, \nonumber\\ &&G=3.14{\kern2mm\rm GeV}^{-2},\quad K=6.1{\kern2mm\rm GeV}^{-5}. \end{eqnarray} We also have a set of additional parameters $c^{\sigma^a(\varphi^a)}_{qq}$ in form factors $f^a_2$. These parameters are defined by masses of excited pseudoscalar mesons, $c_{uu}^{\pi,a_0}=1.44$, $c_{uu}^{\eta,\eta',\sigma,f_0}=1.5$, $c_{us}^{K,K_{0}^{*}}=1.59$, $c_{ss}^{\eta,\eta',\sigma,f_0}=1.66$. The slope parameters $d_{qq}$ are fixed by special conditions satisfying the standard gap equation, $d_{uu}=-1.78 {\kern2mm\rm GeV}^{-2}$, $d_{us}=-1.76 {\kern2mm\rm GeV}^{-2}$, $d_{ss}=-1.73 {\kern2mm\rm GeV}^{-2}$ (see \cite{volk97}). Using these parmeters, we obtain masses of pseudoscalar and scalar mesons which are listed in Table \ref{masses} together with experimental values. \begin{table} \caption{The model masses of mesons, MeV} \label{masses} $$ \begin{array}{||l|c|c|c|c||} \hline \hline & GR & EXC & GR(Exp.) &EXC(Exp.) \\ \hline M_{\sigma} & 530 & 1330 &400-1200 &1200-1500 \\ M_{f_0} & 1070 & 1600 &980\pm10 &1712\pm5 \\ M_{a_0} & 830 & 1500 &983.4\pm0.9 &1474\pm19 \\ M_{\pi} & 140 & 1300 &139.56995\pm0.00035 &1300\pm100 \\ M_{K} & 490 & 1300 &497.672\pm0.031 &1460(?) \\ M_{K_{0}^{*}} & 960 & 1500 &- &1429\pm12 \\ M_{\eta} & 520 & 1280 &547.30\pm0.12 &1297.8\pm2.8 \\ M_{\eta'} & 910 & 1470 &957.78\pm0.14 &1440-1470 \\ \hline \hline \end{array} $$ \end{table} From our calculations we come to the following interpretation of $f_0(1370)$, $f_J(1710)$ and $a_0(1470)$ mesons: we consider them as the first radial excitations of the ground states $f_0(400-1200)$, $f_0(980)$ and $a_0(980)$. Meanwhile, the meson $f_0(1500)$ is much likely a glueball. The strong decays which we consider in the next section substantiate our point of view. \section{Strong decays of the scalar mesons} The ground and excited states of scalar mesons $f_0$, $a_0$ decay mostly into pairs of pseudoscalar mesons. In the framework of a quark model and in the leading order of $1/N_c$ expansion, the processes are described by triangle quark diagrams (see Fig.1). Before we start to calculate the amplitudes, corresponding to these diagrams, we introduce, for convenience, Yukawa coupling constants which naturally appear after the renormalization (\ref{renorm}) of meson fields: \begin{eqnarray} &&g_{\sigma_u}\equiv \left.g_{\sigma^a}\right|_{a=1,2,3,8}=[4I_2(m_u)]^{-1/2}, \quad g_{K_{0}^{*}}\equiv \left.g_{\sigma^a}\right|_{a=4,5,6,7}=[4I_2(m_u,m_s)]^{-1/2}, \nonumber\\ &&g_{\sigma_s}\equiv g_{\sigma^9}=[4I_2(m_s)]^{-1/2}, \quad g_{\varphi^a}=Z^{-1/2}g_{\sigma^a} \nonumber\\ &&g_{\pi}\equiv\left. g_{\varphi^a}\right|_{a=1,2,3}, \quad g_{K}\equiv \left.g_{\varphi^a}\right|_{a=4,5,6,7}, \quad g_{\varphi_u}\equiv g_{\varphi^8}, \quad g_{\varphi_s}\equiv g_{\varphi^9} \label{g} \end{eqnarray} \begin{eqnarray} &&\hat g_{\sigma_u}\equiv \left.\hat g_{\sigma^a}\right|_{a=1,2,3,8}=[4I_2^{ff}(m_u)]^{-1/2}, \;\, \hat g_{K_{0}^{*}}\equiv \left.\hat g_{\sigma^a}\right|_{a=4,5,6,7}=[4I_2^{ff}(m_u,m_s)]^{-1/2}, \nonumber\\ &&\hat g_{\sigma_s}\equiv \hat g_{\sigma^9}=[4I_2^{ff}(m_s)]^{-1/2}, \quad \hat g_{\varphi^a}=\hat g_{\sigma^a} \nonumber\\ &&\hat g_{\pi}\equiv\left.\hat g_{\varphi^a}\right|_{a=1,2,3}, \;\, \hat g_{K}\equiv \left.\hat g_{\varphi^a}\right|_{a=4,5,6,7}, \;\, \hat g_{\varphi_u}\equiv \hat g_{\varphi^8}, \;\, \hat g_{\varphi_s}\equiv \hat g_{\varphi^9} \label{gX} \end{eqnarray} They can easily be related to $Z^a_{\sigma(\varphi),i}$ introduced in the beginning of our paper. Thus, the one-loop contribution to the effective Lagrangian can be rewritten in terms of the renormalized fields: \begin{eqnarray} {\cal L}_{\rm 1-loop}(\sigma,\varphi)&=& i N_c {\rm Tr}\ln\left[1+ \frac{1}{i\partial\hspace{-.5em}/\hspace{.15em}-m}\sum_{a=1}^9 \tau_a\left[ g_{\sigma^a}\sigma^a_1+i\gamma_5 g_{\varphi^a}\varphi^a_1+\right.\right.\nonumber\\ &&\left.(\hat g_{\sigma^a}\sigma^a_2+i\gamma_5 \hat g_{\varphi^a}\varphi^a_2)f_a\right] \Biggr] \label{bosLag1} \end{eqnarray} All amplitudes that describe processes of the type $\sigma\to\varphi_1\varphi_2$ can be divided into two parts: \begin{eqnarray} T_{\sigma\to\varphi_1\varphi_2}&=& C\left(-\frac{i N_c}{(2\pi)^4}\right) \int_{\Lambda_3}d^4 k \frac{{\rm Tr}[(m+\Slash{k}+\Slash{p}_1)\gamma_5 (m+\Slash{k})\gamma_5(m+\Slash{k}-\Slash{p}_2)]}{ (m^2-k^2)(m^2-(k+p_1)^2)(m^2-(k-p_2)^2)}\nonumber\\ &=& 4mC\left(-\frac{i N_c}{(2\pi)^4}\right) \int_{\Lambda_3}d^4k \frac{\left[1-\frac{p_1\cdot p_2}{m^2-k^2}\right]}{(m^2-(k+p_1)^2)(m^2-(k-p_2)^2)} \nonumber\\ &=&4 m C [I_2(m,p_1,p_2)-p_1\cdot p_2 I_3(m,p_1,p_2)]=T^{(1)}+T^{(2)} \label{T} \end{eqnarray} here $C=4 g_{\sigma} g_{\varphi_1}g_{\varphi_2}$ and $p_1$, $p_2$ are momenta of the pseudoscalar mesons. Using (\ref{g}) and (\ref{gX}), we rewrite the amplitude $T_{\sigma\to\varphi_1\varphi_2}$ in another form \begin{eqnarray} &&T_{\sigma\to\varphi_1\varphi_2}\approx 4mZ^{-1/2}g_{\varphi_1} \left[1-p_1\cdot p_2 \frac{I_3(m)}{I_2(m)}\right],\label{T1}\\ &&p_1\cdot p_2 =\frac12(M_\sigma^2-M_{\varphi_1}^2-M_{\varphi_2}^2). \end{eqnarray} We assumed here that the ratio of $I_3$ to $I_2$ slowly changes with momentum in comparison with factor $p_1\cdot p_2$, therefore, we ignore their momentum dependence in (\ref{T1}). With this assumption we are going to obtain just a qualitative picture for decays of the excited scalar mesons. In eqs. (\ref{T}) and (\ref{T1}) we omitted the contributions from the diagrams which include form factors in vertices. The whole set of diagrams consists of those containing zero, one, two and three form factors. To obtain the complete amplitude, one must sum up all contributions. After these general comments, let us consider the decays of $ a_0(1450)$, $f_0(1370)$ and $f_J(1710)$. First, we estimate the decay width of the process $\hat a_0\to\eta\pi$, taking the mixing coefficients from Table~\ref{mixingTable} and~\ref{isoscMixTab} (see Appendix B for the details). The result is \begin{equation} T^{(1)}_{\hat a_0\to\eta\pi}\approx0.2{\kern2mm\rm GeV}, \end{equation} \begin{equation} T^{(2)}_{\hat a_0\to\eta\pi}\approx3.5 {\kern2mm\rm GeV}, \end{equation} \begin{equation} \Gamma_{\hat a_0\to\eta\pi}\approx 160 {\kern2mm\rm MeV}. \end{equation} From this calculation one can see that $T^{(1)}\ll T^{(2)}$ and the amplitude is dominated by its second part, $T^{(2)}$, which is momentum dependent. The first part is small because the diagrams with different numbers of form factors cancel each other. As a consequence, in all processes where an excited scalar meson decays into a pair of ground pseudoscalar states, the second part of the amplitude defines the rate of the process. For the decay $\hat a_0\to\pi\eta'$ we obtain the amplitudes \begin{equation} T^{(1)}_{\hat a_0\to\pi\eta'}\approx0.8 {\kern2mm\rm GeV}, \end{equation} \begin{equation} T^{(2)}_{\hat a_0\to\pi\eta'}\approx3 {\kern2mm\rm GeV}, \end{equation} and the decay width \begin{equation} \Gamma_{\hat a_0\to\pi\eta'}\approx36 {\kern2mm\rm MeV}. \end{equation} The decay of $\hat a_0$ into kaons is described by the amplitudes $T_{\hat a_0\to K^+K^-}$ and $T_{\hat a_0\to \bar K^0K^0}$ which, in accordance with our scheme, can again be divided into two parts: $T^{(1)}$ and $T^{(2)}$ (see Appendix B for details): \begin{equation} T_{\hat a_0\to K^+K^-}^{(1)}\approx 0.2{\kern2mm\rm GeV}, \end{equation} \begin{equation} T_{\hat a_0\to K^+K^-}^{(2)}\approx 2.1{\kern2mm\rm GeV}. \end{equation} and the decay width is \begin{equation} \Gamma_{\hat a_0\to KK}=\Gamma_{\hat a_0\to K^+K^-}+\Gamma_{\hat a_0\to \bar K^0K^0}\approx 100{\kern2mm\rm MeV}. \end{equation} Qualitatively, our results do not contradict the experimental data. \begin{equation} \Gamma^{\rm tot}_{\hat a_0}=265\pm13 {\kern2mm\rm MeV},\quad BR(\hat a_0\to KK):BR(\hat a_0\to\pi\eta)= 0.88\pm0.23. \end{equation} The decay widths of radial excitations of scalar isoscalar mesons are estimated in the same way as it was shown above. We obtain: \begin{equation} \Gamma_{\hat\sigma\to\pi\pi}=\left\{ \begin{array}{l} 550 {\kern2mm\rm MeV} (M_{\hat\sigma}=1.3 {\kern2mm\rm GeV}) \\ 460 {\kern2mm\rm MeV} (M_{\hat\sigma}=1.25 {\kern2mm\rm GeV}), \end{array} \right. \end{equation} \begin{equation} \Gamma_{\hat\sigma\to\eta\eta}=\left\{ \begin{array}{l} 24 {\kern2mm\rm MeV} (M_{\sigma}=1.3 {\kern2mm\rm GeV}) \\ 15 {\kern2mm\rm MeV} (M_{\sigma}=1.25 {\kern2mm\rm GeV}), \end{array} \right. \end{equation} \begin{equation} \Gamma_{\hat\sigma\to\sigma\sigma}=\left\{ \begin{array}{l} 6 {\kern2mm\rm MeV} (M_{\sigma}=1.3 {\kern2mm\rm GeV}) \\ 5 {\kern2mm\rm MeV} (M_{\sigma}=1.25 {\kern2mm\rm GeV}), \end{array} \right. \end{equation} \begin{equation} \Gamma_{\hat\sigma\to KK}\sim 5 {\kern2mm\rm MeV}, \end{equation} \begin{equation} \begin{array}{lclclcl} \Gamma_{f_0(1710)\to 2\pi}& =& 3{\kern2mm\rm MeV}, &\quad & \Gamma_{f_0(1500)\to 2\pi}&=& 3{\kern2mm\rm MeV},\\ \Gamma_{f_0(1710)\to 2\eta}& =& 40{\kern2mm\rm MeV}, &\quad & \Gamma_{f_0(1500)\to 2\eta}&=& 20{\kern2mm\rm MeV},\\ \Gamma_{f_0(1710)\to \eta\eta'}& =& 42{\kern2mm\rm MeV}, &\quad & \Gamma_{f_0(1500)\to \eta\eta'}&=& 10{\kern2mm\rm MeV},\\ \Gamma_{f_0(1710)\to KK}& =& 24{\kern2mm\rm MeV}, &\quad & \Gamma_{f_0(1500)\to KK}&=& 20{\kern2mm\rm MeV}. \end{array} \end{equation} The decays of $f_0(1500)$ and $f_0(1710)$ to $\sigma\sigma$ are negligibly small, so we disregard them. Here we desplayed our estimates for both $f_J(1710)$ and $f_0(1500)$ resonances. Comparing them will allow us to decide which one to consider as the first radial excitation of $f_0(980)$ and which a glueball. From the experimental data: \begin{equation} \Gamma^{\rm tot}_{\sigma'}=200 - 500 {\kern2mm\rm MeV}, \quad \Gamma^{\rm tot}_{f_0(1710)}=133\pm 14 {\kern2mm\rm MeV}, \quad \Gamma^{\rm tot}_{f_0(1500)}=112\pm 10 {\kern2mm\rm MeV} \end{equation} we can see that in the case of $f_0(1500)$ being a $\bar qq$ state there is deficit in the decay widths whereas for $f_J(1710)$ the result is close to experiment. From this we conclude that the meson $f_J(1710)$ is a radially excited partner for $f_0(980)$ and the meson state $f_0(1370)$ is the first radial excitation of $f_0(400-1200)$. As to the state $f_0(1500)$, it is mostly a glueball which significantly contributes to the decay width. For the decay widths of ground scalar states the situation is opposite to that for excited states. Indeed, the parts $T^{(1)}$ and $T^{(2)}$ of the amplitude $T$ are of the same order of magnitude and different in sign, thus, cancelling each other. The values for decay widths of processes $\sigma\to\pi\pi$, $f_0(980)\to\pi\pi$ and $a_0(980)\to\pi\eta$ turn out to be too small in our approximation: \begin{equation} \Gamma_{\sigma\to\pi\pi}\sim 350 {\kern2mm\rm MeV},\quad \Gamma_{f_0(980)\to\pi\pi}\sim 25{\kern2mm\rm MeV},\quad \Gamma_{a_0(980)\to\pi\eta}\sim 1{\kern2mm\rm MeV}, \end{equation} whereas the experiment gives us \begin{eqnarray} &&\Gamma_{\sigma\to\pi\pi}\sim 600-1000 {\kern2mm\rm MeV},\quad \Gamma_{f_0(980)\to\pi\pi}\sim 40-100{\kern2mm\rm MeV},\nonumber\\ &&\Gamma_{a_0(980)\to\pi\eta}\sim 50-100{\kern2mm\rm MeV}. \end{eqnarray} Therefore, to describe correctly strong decays of ground states of scalar mesons, it is necessary to calculate the parts $T^{(1)}$ of the amplitudes more accurately, taking into account their momentum dependence. Indeed, from our previous paper \cite{conf} we know that taking into account the quark confinement and the momentum dependence of $I_2(m,p_1,p_2)$ one can find that $\sigma$ decays into $\pi\pi$ with the width \begin{equation} \Gamma_{\sigma\to\pi\pi}\approx 700{\kern2mm\rm MeV}. \end{equation} For the decay $a_0\to\eta\pi$, the result can increase more than by an order. A proper description of the decay $a_0(980)\to\pi\eta$ can possibly be obtained from the four-quark interpretation of $a_0(980)$ state \cite{achasov}, and the four-quark component may dominate the process; however, this is beyond our model. For a more careful description of the decay $f_0(980)\to\pi\pi$ one should take into account also the mixing with the glueball state $f_0(1500)$, which we are going to do in our further work. \section{Discussion and Conclusion} Our calculations have shown that we can interpret the scalar states $f_0(1370)$, $a_0(1450)$ and $f_0(1710)$ as the first radial excitations of $f_0(400-1200)$, $a_0(980)$ and $f_0(980)$. We estimated their masses and the widths of main decays in the framework of a nonlocal chiral quark model. We would like to emphasize that we have not used additional parameters except those necessary to fix the mass spectrum of pseudoscalar mesons. We used the same form factors both for the scalar and pseudoscalar mesons, which is a requirement of the global chiral symmetry. We assumed that the state $f_0(1500)$ is a glueball, and its probable mixing with $f_0(980)$, $f_0(1370)$ and $f_J(1710)$ may provide us with a more correct description of the masses of these states% \footnote{ Our estimates for the masses of $f_0$ and $\hat f_0$: $M_{f_0}=1070{\kern2mm\rm MeV}$ and $M_{\hat f_0}=1600{\kern2mm\rm MeV}$ are expected to shift to $M_{f_0}=980{\kern2mm\rm MeV}$ and $M_{\hat f_0}=1710{\kern2mm\rm MeV}$ after mixing with the glueball $f_0(1500)$. }% (see Table~\ref{masses}). We are going to consider this problem in a subsequent publication. More complicated situation takes place for the ground state $a_0(980)$. In the framework of our quark-antiquark model, we have a mass deficit for this meson, $830 {\kern2mm\rm MeV}$ instead of $980{\kern2mm\rm MeV}$ and a small decay width for $a_0\to\pi\eta$. We suspect that this drawback is caused by four-quark component in this state which we did not take into account \cite{achasov}. There is also some mystery about the strange scalar meson $K_{0}^{*}$. Its experimental mass is large enough, $M_{K_{0}^{*}}=1430{\kern2mm\rm MeV}$ and the width is $287\pm23{\kern2mm\rm MeV}$. In our model, there are two strange scalars, $K_{0}^{*}(930)$ with a large width and $K_{0}^{*}(1500)$ with the width of the decay $\Gamma_{K_{0}^{*}\to K\pi}\sim300{\kern2mm\rm MeV}$. Thus, together with \cite{ishida} we suppose it is possible for a wide strange resonance, $K_{0}^{*}(930)$ to exist in nature still missed in detectors as the ground state whereas the resonance $K_{0}^{*}(1430)$ is its radial excitation mostly decaying into $K\pi$. Our model gives for the excited meson $K_{0}^{*}$: $M_{\hatK_{0}^{*}}\approx 1500{\kern2mm\rm MeV}$ and $\Gamma_{\hatK_{0}^{*}\to K\pi}\approx 300{\kern2mm\rm MeV}$. In future we are going to take into account the presence of glueball states and to develop a model with quark confinement for the description of heavy mesons. We will also consider the decays of the excited $\eta$ and $\eta'$ mesons. \section*{Acknowledgment} We are very grateful to Prof. S.B.~Gerasimov for useful discussion. This work has been supported by RFBR Grant N 98-02-16135 and by Heisenberg--Landau Program 1999.
\section{Introduction}\label{intro} The interplay between chaos, quantum mechanics and dissipation is rather complex and the subject of strong current research activities \cite{Dittrich98,Habib98,Cohen97,Dittrich90,Graham86,Graham85}. The definition of chaos in classical mechanics via exponentially fast spreading trajectories can not be applied to quantum mechanical systems, since the notion of a trajectory does not exist in quantum mechanics. On a quantum mechanical level chaos manifests itself in the statistical properties of the eigenenergies and eigenfunctions. In the case of Hamiltonian systems the eigenenergies and eigenfunctions obey the universal statistics of large random hermitian matrices restricted only by general symmetry requirements like invariance under time and spin reversal \cite{Bohigas84,Berry83}. While no rigorous proof of this conjecture is known yet, overwhelming numerical and experimental evidence has been accumulated \cite{Guhr98,Reichl92,Haake91}.\\ Dissipation has at least two very important effects. The classical dynamics is altered profoundly. It is no longer restricted to a shell of constant energy in phase space, but phase space volume shrinks if no external source compensates for the energy dissipated. In a chaotic system with external driving dissipation typically leads to a strange attractor in phase space, i.e.~a multi-fractal structure that is invariant under the dynamics and which has a dimension strictly smaller than the dimension of the phase space. The second effect is of quantum mechanical nature and more subtle: Dissipation destroys very efficiently the quantum mechanical phase information. This typically happens on time scales much shorter than classical ones and even with very tiny amounts of dissipation \cite{Zurek81,Giulini96,DBraun98.2}. Therefore the system behaves more classically, and one might expect to find classical manifestations of chaos again. It was indeed shown in a variety of examples that appropriate quantum mechanical counterparts of the classical phase space density (like Husimi functions or Wigner distributions) approach a smeared out version of the strange attractor \cite{Dittrich90,Peplowski91,Kolovsky96}. At the same time one might ask whether spectral properties approach their classical counterparts as well. This paper shows that for certain spectral properties the answer is ``YES'', even though the structure of the spectrum can be very different in both cases. \\ My analysis is based on a recently discovered trace formula for dissipative systems which, in the spirit of Gutzwiller's celebrated formula \cite{Gutz70,Gutz71}, expresses traces of the propagator of the density matrix in terms of classical periodic orbits \cite{DBraun98}. I show in section \ref{sectrace} that to lowest order asymptotic expansion in $\hbar$ the traces agree with the traces of the classical Frobenius--Perron propagator of the phase space density \cite{Cvitanovic91}.\\ In the next section I briefly review basic properties of dissipative quantum maps, semiclassical theory and the trace formula. In section \ref{secnum} I apply the trace formula to a dissipative kicked top and compare with numerical results for finite $\hbar$. The main results are summarized in section \ref{summary}. \section{Dissipative Quantum Maps} \subsection{General remarks} Dissipation is introduced on a quantum mechanical level most rigorously by the so--called Hamiltonian embedding \cite{Weiss93}. The system of interest is considered as part of a larger system including the ``environment'' to which energy can be dissipated. The total system is assumed to be closed so that it is adequately described by a Schr\"odinger equation. The degrees of freedom of the ``environment'' remain unobserved. The system at interest is described by a density matrix in which the environmental degrees of freedom have been traced out, usually termed the reduced density matrix $\rho(t)$. \\ Dissipative quantum maps are maps of the reduced density matrix from a time $t$ to a time $t+T$: $\rho(t+T)=P\rho(t)$. They are analogous to non--dissipative quantum maps, where the state vector of the system is mapped with a unitary Floquet matrix $F$, $|\psi(t+T)\rangle =F|\psi(t)\rangle$. In the dissipative case $P$ is not a unitary operator and therefore has eigenvalues inside the unit circle.\\ Maps, dissipative or not, are a natural description of a time evolution if an external driving of the system is periodic in time with period $T$. They give a stroboscopic picture which suffices if the evolution during one period is of no interest. Systems that are periodically driven are capable of chaos even if they have only one degree of freedom. I will restrict myself in the following to such cases. The maps that I consider are particularly simple in the sense that the dissipation is well separated from a remaining purely unitary evolution where the latter by itself is capable of chaos. The unitary part will be described by a Floquet matrix $F$ acting on the state vector, so that the unitary evolution takes the density matrix from $\rho(t)$ to $\rho'(t)=F\rho(t)F^\dagger$. After the unitary part a dissipative step follows which I will describe by a propagator $D$. It takes the density matrix from $\rho'(t)$ to $\rho(t+T)=D\rho'(t)$. The total map therefore reads \begin{equation} \label{map} \rho(t+T)=D(F\rho(t) F^\dagger)\equiv P\rho(t)\,. \end{equation} Such a separation into two parts is not purely academic. A most obvious realization of (\ref{map}) is given when a Hamiltonian $H(t)$ leading to the unitary evolution and the coupling to the environment can be turned on and off alternatively. This should be realizable for instance with atoms flying through a series of cavities where in each cavity either the unitary evolution or the dissipation is realized. Another example might be a billiard, in which the particle only dissipates energy when hitting the walls. But even if the dissipation cannot be turned off, the map (\ref{map}) may still be a good description. For instance, if the dissipation is weak and if the entire unitary evolution takes place during a very short time, dissipation may be negligible during that time. This is the case if the entire unitary evolution is due to a periodic kicking. The dissipation can then be considered as a relaxation process between two successive kicking events. Finally, a formal reason for such a separation can be given when the generators for the unitary evolution and the dissipation commute. These ideas might become clearer with a particular model. Let me therefore introduce as prime model system a dissipative kicked top.\\ \subsection{A dissipative kicked top} The dynamical variables of a top \cite{kickedtop,Haake91} are the three components $J_{x,y,z}$ of an angular momentum ${\bf J}$. I will only consider dynamics (including the dissipative ones) which conserve the absolute value of ${\bf J}$, ${\bf J}^2=j(j+1)=const$. In the classical limit (formally attained by letting the quantum number $j$ approach infinity) the surface of the unit sphere $\lim_{j\rightarrow\infty} ({\bf J}/j)^2=1$ becomes the phase space, such that one confronts but a single degree of freedom. Convenient phase space coordinates are \begin{equation} \label{pq} \mu\equiv J_z/J=\cos\theta=p\mbox{ and } \varphi=q, \end{equation} where the polar and azimuthal angles $\theta$ and $\varphi$ define the orientation of the angular momentum vector with respect to the $J_x$, $J_y$ and $J_z$ axes. The parameter $J$ is defined as $J=j+1/2$ and allows for more convenient expressions of most semiclassical quantities. Due to the conservation of ${\bf J}^2$ the Hilbert space decomposes into $(2j+1)$ dimensional subspaces. The quantum dynamics is confined to one of these according to the initial conditions. The semiclassical limit is characterized by large values of the quantum number $j$ which can be integer or half integer. Since the classical phase space contains $(2j+1)$ states, Planck's constant may be thought of as represented by $1/J$. \\ Consider a unitary evolution generated by the Floquet matrix \begin{equation} \label{F} F={\rm e}^{-{\rm i} \frac{k}{2J}J_z^2}{\rm e}^{-{\rm i} \beta J_y}\,. \end{equation} The corresponding classical motion first rotates the angular momentum by an angle $\beta$ about the $y$-axis and then subjects it to a torsion about the $z$-axis. The latter may be considered as a non--linear rotation with a rotation angle given by the $J_z$ component of ${\bf J}$. The dynamics is known to become strongly chaotic for sufficiently large values of $k$ and $\beta$, whereas either $k=0$ or $\beta=0$ lead to integrable motion \cite{Haake91}. For a physical realization of this dynamics it might be best to think of ${\bf J}$ as a Bloch vector describing the collective excitations of two--level atoms, as one is used to in quantum optics. The rotation can be brought about by a laser pulse of suitably chosen length and intensity, and the torsion by a cavity that is strongly off resonance from the atomic transition frequency \cite{Agarwal97}. The Floquet matrix (\ref{F}) has also been realized in experiments with magnetic crystallites with an easy plane of magnetization \cite{Waldner85}. Our model dissipation is defined in continuous time $\tau$ by the Markovian master equation \begin{equation} \label{rhotd} \frac{d}{d\tau}\rho(\tau)=\frac{1}{2J}([J_-,\rho(\tau) J_+]+[J_-\rho(\tau), J_+])\equiv \Lambda\rho(\tau)\,, \end{equation} where the linear operator $\Lambda$ is defined by this equation as generator of the dissipative motion. Equation (\ref{rhotd}) is well-known to describe certain superradiance experiments, where a large number of two--level atoms in a cavity of bad quality radiate collectively \cite{Bonifacio71.1,Haroche82}. The angular momentum operator ${\bf J}$ is then again the Bloch vector of the collective excitation and the $J_+,J_-$ are raising and lowering operators, $J_\pm=J_x\pm{\rm i} J_y$. One easily verifies that (\ref{rhotd}) conserves the skewness in the $J_z$ basis ($J_z|m\rangle =m|m\rangle$), i.e.~matrix elements $\langle m_1 |\frac{d}{d\tau}\rho|m_2\rangle$ with a given skewness $J\eta=m_1-m_2$ depend only on matrix elements with the same skewness. Eq.(\ref{rhotd}) is formally solved by $\rho(\tau)=\exp(\Lambda\tau)\rho(0)$ for any initial density matrix $\rho(0)$ and this defines the dissipative propagator \begin{equation} \label{D} D(\tau)=\exp(\Lambda\tau)\,. \end{equation} Explicit forms of $D$ can be found in \cite{Bonifacio71.1,PBraun98.1,PBraun98.2}. The skewness $\eta$ only enters as a parameter in $D$. The classical limit gives the simple picture of the Bloch vector creeping towards the south pole $\theta=\pi$ as an over-damped pendulum, according to the equations of motion \begin{equation} \label{eomdiss} \frac{d}{d\tau}\theta=\sin\theta\mbox{, }\frac{d}{d\tau}\varphi=0\,. \end{equation} Classically the azimuth $\varphi$ is therefore conserved. Eq.(\ref{eomdiss}) also shows that $\tau$ is the time in units of the classical time scale. In the following it will be set equal to the time between two unitary steps. \\ The Floquet matrix (\ref{F}) is usually generated by the Hamiltonian \begin{equation} \label{Hoft} H(t)=\hbar\left(\frac{k}{2JT}J_z^2+\beta J_y\sum_{n=-\infty}^\infty\delta(t-nT)\right); \end{equation} it describes the evolution from immediately before a kick to immediately before the next kick. The generator (\ref{rhotd}) for the dissipation does not commute with $H(t)$. In order to obtain the map (\ref{map}) one should replace in (\ref{Hoft}) $H_0=\frac{k}{2JT}J_z^2$ by $\frac{k}{2JT_1}J_z^2$ and switch on $H(t)$ only for a time $T_1<T$ during each period $T$, whereas the dissipation acts during the rest of the time $\tau=T-T_1$. Alternatively, when $H(t)$ and $\Lambda$ act permanently one may go to an interaction representation by $\rho(t)= \exp(-\frac{{\rm i}}{\hbar}H_0t)\tilde{\rho}(t)\exp(\frac{{\rm i}}{\hbar}H_0t)$. In the $J_z$ representation this leads only to phase factors in the master equation (\ref{rhotd}) which can be easily incorporated in $D$ and which vanish moreover for diagonal elements. Let us assume in the following that either has been done and use (\ref{map}) with $F$ and $D$ given by (\ref{F}), (\ref{rhotd}), and (\ref{D}) as a starting point with $\tau$ as fixed parameter that measures the relaxation time between two unitary evolution and thus the dissipation strength. \subsection{The trace formula} In 1970 Gutzwiller published a trace formula for Hamiltonian flows that has become a center piece of subsequent studies of quantum chaotic systems \cite{Gutz70,Gutz71}. A corresponding formula was obtained later for non-dissipative quantum maps by Tabor \cite{Tabor83}. Assuming the existence of a corresponding classical map ${\bf y}={\bf f}({\bf x})$ of phase on itself (${\bf x}=(p',q')$ are the old, ${\bf y}=(p,q)$ the new phase space coordinates), both formulae express a spectral property of the quantum mechanical propagator as a sum over periodic orbits of ${\bf f}$. Each periodic orbit contributes a weighted phase factor, where the weight depends on the stability matrix $M$ of the orbit and the phase is basically given by the classical action $S$ in units of $\hbar$. Tabor's formula aims at traces of the Floquet matrix $F$, \begin{equation} \label{trFN} {\rm tr} F^N=\sum_{p.p.}\frac{{\rm e}^{{\rm i}(\frac{ S}{\hbar}-\frac{\pi}{2}\nu)}}{|2-{\rm tr} M|^{1/2}}\,. \end{equation} I have written the sum over periodic orbits as a sum over periodic points ($p.p.$) of the $N$ times iterated map ${\bf f}^N$; the integer $\nu$ (the so--called Maslov index) counts the number of caustics along the orbit. All quantities have to be evaluated on the periodic points. The squared modulus of ${\rm tr} F^N$ has in the unitary case an interpretation as (discrete time) form factor of spectral correlations.\\ In \cite{DBraun98} a corresponding trace formula for dissipative quantum maps of the form (\ref{map}) was derived. It is based on semiclassical approximations for both $F$ and $D$. The semiclassical approximation of $F$ has the general form of a van Vleck propagator \cite{PBraun96,vanVleck28}; a corresponding semiclassical approximation for $D$ was obtained in \cite{PBraun98.2}. A WKB ansatz lead to a fictitious Hamiltonian system which depends on the skewness as a parameter. Its trajectories connect initial and final points specified by the arguments of $D$. Much as in the unitary case, an action $R$ is accumulated along the trajectories; it has the usual generating properties of an action. Based only on the general van Vleck forms of $F$ and $D$ and the generating properties of the actions $S$ and $R$ we derived the trace formula \begin{equation} \label{trPN} {\rm tr} P^N=\sum_{p.p.}\frac{{\rm e}^{\sum_{i=1}^N JR_i}}{\left|{\rm tr} \prod_{i=N}^1M_d^{(i)}-{\rm tr} M\right|}\mbox{, }\quad\quad N=1,2,\ldots\,. \end{equation} The sum is over all periodic points of the $N$--times iterated dissipative classical map; the $R_i$ are the actions of the fictitious Hamiltonian system for vanishing skewness accumulated during the $i$th dissipative step. The denominator contains the stability matrices $M_d^{(i)}$ for the $i$th dissipative step and $M$ for the entire map ${\bf f}^N$. The matrix $M_d^{(i)}$ with index $i=N$ is at the left of the product. Eq.(\ref{trPN}) is a leading order asymptotic expansion in $1/J$ for propagators $P$ of the type (\ref{map}). The following restrictions apply: \begin{itemize} \item The phase space is two dimensional. \item The classical limit of the dissipative part of the map conserves one phase space coordinate (the azimuthal coordinate $\varphi$ for the dissipation described by (\ref{rhotd})). \item The propagator $D$ for the dissipative part conserves the skewness $\eta$ of the density matrix in a suitably chosen basis and $D$ has a single maximum as a function of $\eta$ at $\eta=0$. As indicated the dissipation (\ref{rhotd}) conserves the skewness in the $J_z$ basis. \item The dissipation exceeds a certain minimum value. It is given by $\tau\gtrsim 1/J$ for the dissipation (\ref{rhotd}) and thus may become infinitesimally small in the classical limit $J\to\infty$. \end{itemize} Eq.(\ref{trPN}) shows that periodic orbits and classical quantities related to them still determine the spectral properties of the quantum system even in the presence of dissipation. The formula will now be studied in more detail. \section{Connection to classical trace formula}\label{sectrace} Remarkable about (\ref{trPN}) is its simplicity. First of all, when propagating a density matrix, one would expect a double sum over periodic points. Indeed, in the dissipation free case one easily shows ${\rm tr} P=|{\rm tr} F|^2$, and ${\rm tr} F$ is given by the Tabor formula (\ref{trFN}) as a simple sum over periodic points \cite{Tabor83}. Out of the double sum, only the ``diagonal parts'' survive. Decoherence induced through dissipation destroys the interference terms between different periodic points. For the diagonal terms the actions $S$ and $-S$ stemming from $F$ and $F^\dagger$ and the phases due to the Morse indices cancel. The square roots in the denominator combine to a power 1. Due to the cancellation of the phase factors the traces (\ref{trPN}) are always real and positive. They fulfill herewith a general requirement for all propagators of density matrices that follows from conservation of positivity of the density matrix. On the other hand one may wonder whether the trace formula should not be an entirely classical formula, if all interference terms are destroyed. This is indeed what I am going to show now. The classical propagator of phase space density is given by $P_{cl}({\bf y},{\bf x})=\delta({\bf y}-{\bf f}({\bf x}))$. In the case where $P_{cl}({\bf y},{\bf x})$ describes the map arising from the evolution during a finite time of an autonomous system, $P_{cl}$ is commonly called the Frobenius--Perron operator. For brevity I use the same name in the present dissipative situation. The trace of the $N$th iteration of $P_{cl}$ is given by \cite{Cvitanovic91} \begin{eqnarray} {\rm tr} P_{cl}^N&=&\sum_{p.o.}\sum_{r=1}^\infty \frac{n_p\delta_{N,n_pr}}{|\det({\bf 1}-M_p^r)|} \label{first}\\ &=&\sum_{p.p.}\frac{1}{|\det({\bf 1}-M)|}\,,\label{second} \end{eqnarray} where the first sum in (\ref{first}) is over all primitive periodic orbits of length $n_p$, $r$ is their repetition number and $M_p$ the stability matrix of the primitive orbit. In (\ref{second}), $p.p.$ labels all periodic points belonging to a periodic orbit of total length $N$, including the repetitions, and $M$ is the stability matrix for the entire orbit. \\ The fact that $M$ in (\ref{trPN}) is a $2\times 2$ matrix leads immediately to $\det({\bf 1}-M)=1+\det M-{\rm tr} M$. Since the map is a periodic succession of unitary evolutions (with stability matrices $M_u^{(i)}$) and dissipative evolutions (with stability matrices $M_d^{(i)}$), $M$ is given by the product $M=\prod_{i=N}^1M_d^{(i)}M_u^{(i)}$. The stability matrices $M_u^{(i)}$ are all unitary so that $\det M_u^{(i)}=1$ for all $i=1\ldots N$ and $\det M=\prod_{i=N}^1\det M_d^{(i)}$. The dissipative process for which (\ref{trPN}) was derived conserves $q$ which means that $M_d^{(i)}$ is diagonal, \begin{equation} \label{md} M_d^{(i)}=\left(\begin{array}{cc}1 &0\\ 0&m_d^{(i)} \end{array} \right)\,. \end{equation} The upper left element is $\frac{\partial q(p',q')}{\partial q'}$, the lower right $\frac{\partial p(p',q')}{\partial p'}$. But then $\det M_d^{(i)}=m_d^{(i)}$, and we find with $\left|{\rm tr} \prod_{i=N}^1 M_d^{(i)}-{\rm tr} M\right|=|1+\det M-{\rm tr} M|= |\det({\bf 1}-M)|$ exactly the denominator in (\ref{second}). \\ The actions $R_i$ are zero on the classical trajectories for the dissipative process (\ref{rhotd}), as one immediately sees by using their explicit form \cite{PBraun98.2}. Their vanishing can be retraced more generally to conservation of probability by the master equation and therefore holds for other master equations of the same structure as well. To see this write (\ref{rhotd}) in the $J_z$ basis and look at the part with vanishing skewness, i.e.~the probabilities $p_m=\langle m|\rho|m\rangle$. We obtain a set of equations \begin{equation} \label{pt} \frac{d}{d\tau}p_m=(g_{m+1}p_{m+1}-g_m p_m)\,, \end{equation} where the specific form of the coefficients $g_m$ is of no further concern. Important is rather that the {\em same} function $g_m$ appears twice. This is sufficient and necessary for the conservation of probability, ${\rm tr} \rho=\sum_{m=-j}^jp_m=1$. On the other hand, had we coefficients $f_m$ and $g_m$ (i.e.~$\dot{p}_m=(g_{m+1}p_{m+1}-f_m p_m)$) we would obtain the action $R$ on the classical trajectory as $JR=\sum_{l=m}^n(\ln(g_l)-\ln(f_l))$ as one easily verifies by writing down the exact Laplace image of $D$ following the lines in \cite{PBraun98.1}. Thus, the action is zero iff probability is conserved. But then {\em the trace formula (\ref{trPN}) is identical to the classical trace formula (\ref{second}).} This result proves that the traces of any finite power of the evolution operator of the quantum mechanical density matrix, are, in the limit of $\hbar\to 0$, exactly given by the corresponding traces of the evolution operator of the classical phase space density, provided a small amount of dissipation is introduced. This is quite surprising since it is clear that even the basic structure of the two spectra can be very different: For all finite Hilbert space dimensions $d=2j+1$ the quantum mechanical propagator $P$ can be represented as a finite $d^2\times d^2$ matrix. Its spectrum is therefore always discrete, regardless of whether the corresponding classical map is chaotic or not. On the other hand, it is known that $P_{cl}$ has necessarily a continuous spectrum if the classical dynamics is mixing \cite{Gaspardbook}. A formal reason why the spectra may differ in spite of the fact that the traces agree to lowest order in $\hbar$ is easily found. In order to construct the entire spectrum of $P$ one needs $d^2=(2j+1)^2$ traces. But already for traces of order $j$ the next order corrections in the asymptotic expansion in $1/J$ that lead to (\ref{trPN}) become comparable to the classical term; and for the highest traces needed (i.e. traces of order $j^2$) the next order in $1/J$ would be even more important than the classical term, so that one may not expect ${\rm tr} P^{j^2}={\rm tr} P_{cl}^{j^2}$ for $j\to\infty$. In other words, if we do {\em not} keep $n$ fixed in the classical limit, $ {\rm tr} P^n={\rm tr} P_{cl}^n$ may not hold for $j\to \infty$ and therefore the spectrum of $P$ can be very different from that of $P_{cl}$. \\ The asymptotic equality of ${\rm tr} P^n$ and ${\rm tr} P_{cl}^n$ strengthens substantially the quite envolved semiclassical derivation of (\ref{trPN}) \cite{DBraun98}. It also sheds light on the question what happens if the dissipation does not conserve the coordinate $q$. We should then not expect (\ref{trPN}) to be valid but presumably replace it with the more general form (\ref{second}). \section{Comparison with numerical results}\label{secnum} The question arises how good the agreement between quantum and classical traces is for finite $J$. To answer this question I have calculated numerically the exact quantum mechanical traces for our dissipative kicked top and compared them with the traces obtained from the trace formula (\ref{trPN}). These results will be presented now. \subsection{The first trace} The quantum mechanical propagator $P$ is most conveniently calculated in the $J_z$ basis, since the torsion part is then already diagonal. The rotation about the $y$--axis leads to a Wigner $d$--function whose values are obtained numerically via a recursion relation as described in \cite{PBraun96}. The propagator for the dissipation is obtained by inverting numerically the exactly known Laplace image \cite{Bonifacio71.1,PBraun98.1}. The total propagator $P$ is a full, complex, non--hermitian, and non--unitary matrix of dimension $(2j+1)^2\times (2j+1)^2$. Since for the first trace the knowledge of the diagonal matrix elements suffices I was able to calculate ${\rm tr} P$ up to $j=80$. Higher traces are most efficiently obtained via diagonalization which limited the numerics to $j\le 40$.\\ The effort for calculating the first classical trace is comparatively small. In all examples considered and even in the presence of a strange attractor, $P_{cl}$ had at most 4 fixed points that could easily be found numerically by a simple Newton--method in two dimensions. For each fixed point the stability matrix is found via the formulae in Appendix \ref{appA} and so the trace is immediately obtained .\\ In Fig.\ref{figtrPk4b2} I show ${\rm tr} P$ for different values of $j$ as a function of $\tau$ and compare with ${\rm tr} P_{cl}$, eq.(\ref{trPN}). The values for torsion strength and rotation angle, $k=4.0$ and $\beta=2.0$ were chosen such that the system is already rather chaotic in the dissipation free case at $\tau=0$; a phase space portrait of many iterations of $P_{cl}$ shows a large chaotic sea and 6 relatively small stable islands. When $\tau$ reaches a value of the order $\tau\simeq 0.5$ a strange attractor appears which rapidly changes its form and dimension when $\tau$ is increased. The attractor shrinks and is pushed more and more towards the south pole, as the angular momentum has more and more time to relax towards the ground state $J_z=-j$ between two kicks. At values of $\tau$ of the order of $\tau\simeq 2.0-3.0$ the attractor degenerates to a strong point attractor close to the south pole which absorbs even remote initial points in very few steps. At even stronger damping the point attractor reaches the south pole asymptotically.\\ Figure \ref{figtrPk4b2} shows that -- with the exception of very small damping -- ${\rm tr} P_{cl}$ reproduces ${\rm tr} P$ perfectly well for all $\tau$, in spite of the strongly changing phase space structure. The agreement extends to smaller $\tau$ with increasing $j$, as is to be expected from the condition of validity of the semiclassical approximation, $\tau\gtrsim 1/J$ \cite{PBraun98.2}. The analysis of the fixed points shows that at $k=4.0$, $\beta=2.0$ always two fixed points exist for $\tau\gtrsim 0.1$. Their $\mu$ component slowly decreases and the lower one converges towards the south pole with increasing $\tau$, where it finally coincides with the point attractor. \\ Fig.\ref{figfixedpk8b2} shows the fixed point structure for a more complicated situation ($k=8.0$, $\beta=2.0$). The dissipation free dynamics at $\tau=0$ is entirely chaotic, no visible phase space structure is left. The above statements about the creation of a strange attractor (see Fig.\ref{figSAk8b2t1}) and its degeneration to a point attractor when $\tau$ is increased apply equally well. In Fig.\ref{figtrPk8b2} I show the first trace as function of $\tau$ for this situation. The classical trace diverges whenever a bifurcation is reached. Such a behavior is well known from the Gutzwiller formula in the unitary case; the reason for the divergence is easily identified as breakdown of the saddle point approximation in the semiclassical derivation of the trace formula. Whereas the quantum mechanical traces for small $j$ (say $j\simeq 10$) seem not to take notice of the bifurcations, they approximate the jumps and divergences better and better when $j$ is increased. At $j=80$ the agreement with the classical trace is already very good between the bifurcations. Remarkable, however, is the fact that there are some values of $\tau$ close to the bifurcations, where all ${\rm tr} P$ curves for different $j$ in the entire $j$ range examined cross. The trace seems to be independent of $j$ at these points, but they nevertheless do not lie on the classical curve. One is reminded of a Gibbs phenomenon, but I do not have any explanation for it. \subsection{Higher Traces} Let us now examine higher traces ${\rm tr} P^N$ for given values of $k$, $\beta$, and $\tau$ as a function of $N$. For large $N$ all higher traces must converge exponentially to $1$, independent of the system parameters. This is due to the fact that $P$ has always one eigenvalue equal to $1$. Its existence follows from elementary probability conservation \cite{Haake91}. The corresponding eigenmode is an invariant density matrix, its classical counterpart the (strange or point) attractor, the fixed points or any linear combinations thereof \cite{Gaspardbook}. All other eigenvalues have an absolute value smaller than $1$ since there is only dissipation and no amplification in the system. Their powers decay to zero as a function of $N$.\\ I will focus on two limiting cases: The case where the basic phase space structure is a point attractor and the case where it is a well extended strange attractor. As explained above a point attractor can always be obtained by sufficiently strong damping. Consider the example $k=4.0$, $\beta=2.0$ and $\tau=4.0$. Fig.\ref{figtrPNk4b2t4} shows that indeed both quantum mechanical and classical result converge rapidly towards $1$, and the agreement is very good even for $j=10$. If one examines the convergence rate one finds that it is slightly $j$-dependent, but rapidly reaches the classical value. It should be noted that the calculation of ${\rm tr} P_{cl}^N$ is enormously simplified here by the fact that with increasing $N$ no additional periodic points arise. The dissipation is so strong that the system is integrable again. In the example given there are only two fixed points, one at $(\mu,\phi)\simeq (-0.3812219, -3.098751)$, a strong point repeller, and one at $(\mu,\phi)\simeq (-0.9984018,-1.444154)$ a strong point attractor, and all periodic points of $P_{cl}^N$ are just repetitions of these two points. \\ The situation is quite different in the case of a strange attractor (Fig.\ref{figtrPNk8b2t1}). The number of periodic points increases exponentially with $N$, as is typical for chaotic systems. This makes the classical calculation of higher traces exceedingly difficult. For $k=8.0$, $\beta=2.0$, and $\tau=1.0$ I was able to calculate ${\rm tr} P_{cl}^N$ reliably up to $N=5$, where about 400 periodic points have to be taken into account. The obtained numerical result for ${\rm tr} P_{cl}^N$ can always be considered as lower bound for the exact result for ${\rm tr} P_{cl}^N$ as long as one can exclude over-counting of fixed points since all terms in the sum (\ref{trPN}) are positive. It is then clear that at $N=5$ the quantum mechanical result at $j=40$ is still more than a factor 3 away from ${\rm tr} P_{cl}^N$, even though for $N=1$ the agreement is very good. The convergence of ${\rm tr} P^N$ to ${\rm tr} P_{cl}^N$ as a function of $j$ becomes obviously worse with increasing $N$. \\ \section{Summary}\label{summary} I have shown for certain dissipative quantum maps that the traces of (iterations of) the propagator of the quantum mechanical density matrix agrees to first order in an asymptotic expansion in $\hbar$ with the traces of the classical Frobenius--Perron propagator of the phase space density if a small amount of dissipation is present. This holds in spite of the fact that the corresponding spectra are very different. I have tested the theory numerically for finite values of $\hbar$ for a dissipative kicked top and have found good agreement in parameter regimes that ranged from very weak to strong dissipation. The phase space structure turned out to be important in the sense that higher quantum mechanical traces agree with very high precision with the classical ones if the phase space is dominated by a point attractor (strong dissipation), whereas the precision is lost for higher traces in the case of an extended strange attractor (weak dissipation). Sufficiently far away from bifurcations the lowest traces always agree very well with their classical counterpart. {\it Acknowledgments:} I gratefully acknowledge fruitful discussions with P.A.Braun, F.Haake, and J.Weber, and hospitality of the ICTP Trieste, where part of this work was done. Numerical computations were partly performed at the John von Neumann--Institute for Computing in J\"ulich. \begin{appendix} \section{Classical maps and their stability matrices}\label{appA} I give here the classical maps for the three components rotation, torsion and dissipation as well as their stability matrices in phase space coordinates. All maps will be written in the notation $(\mu,\phi)\longrightarrow (\nu,\psi)$, i.e.~$\mu$ and $\nu$ stand for the initial and final momentum, $\phi$ and $\psi$ for the initial and final (azimuthal) coordinate. The latter is defined in the interval from $-\pi$ to $\pi$. The stability matrices will be arranged as \begin{equation} \label{Mgen} M=\left( \begin{array}{cc} \frac{\partial \psi}{\partial \phi}&\frac{\partial \nu}{\partial \phi}\\ \frac{\partial \psi}{\partial \mu}&\frac{\partial \nu}{\partial \mu}\\ \end{array} \right)\,. \end{equation} {\em a. Rotation by an angle $\beta$ about $y$--axis}\\ The map reads \begin{eqnarray} \nu&=&\mu\cos\beta-\sqrt{1-\mu^2}\sin\beta\cos\phi\\ \psi&=&\Big(\arcsin(\sqrt{\frac{1-\mu^2}{1-\nu^2}}\sin\phi)\theta(x')+\\ &&({\rm sign}(\phi)\pi-\arcsin(\sqrt{\frac{1-\mu^2}{1-\nu^2}}\sin\phi) \theta(-x')\Big){\rm mod} 2\pi\\ x'&=&\sqrt{1-\mu^2}\cos\phi\,\cos\beta+\mu\sin\beta\,, \end{eqnarray} where $x'$ is the $x$ component of the angular momentum after rotation, $\theta(x)$ the Heaviside theta--function, and ${\rm sign}(x)$ denotes the sign function. \\ The stability matrix connected with this map is \begin{equation} \label{Mr} M_r=\left( \begin{array}{cc} \sqrt{1-\mu^2}\left(\frac{\cos\phi}{\sqrt{1-\nu^2}\cos\psi}+\frac{\nu\sin\phi\tan\psi\sin\beta}{1-\nu^2}\right)& \,\,\sqrt{1-\mu^2}\sin\phi\sin\beta\\ \frac{\nu\sin\psi(\sqrt{1-\mu^2}\cos\beta+\mu\cos\phi\sin\beta)}{\sqrt{1-\mu^2}(1-\nu^2)\cos\psi}-\frac{\mu\sin\phi}{\sqrt{(1-\nu^2)(1-\mu^2)}\cos\psi} &\cos\beta+\frac{\mu\cos\phi\sin\beta}{\sqrt{1-\mu^2}} \end{array} \right)\,. \end{equation} {\em b. Torsion about $z$--axis}\\ Map and stability matrix are given by \begin{eqnarray} \nu&=&\mu\\ \psi&=&(\phi+k\mu){\rm mod} 2\pi\\ M_t&=&\left( \begin{array}{cc} 1&0\\ k &1\\ \end{array} \right)\label{Mt}\,. \end{eqnarray} {\em c. Dissipation}\\ The dissipation conserves the angle $\phi$, and the stability matrix is diagonal: \begin{eqnarray} \nu&=&\frac{\mu-\tanh\tau}{1-\mu\tanh\tau}\\ \psi&=&\phi\\ M_d&=&\left( \begin{array}{cc} 1&0\\ 0&\frac{1-(\tanh\tau)^2}{(1-\mu\tanh\tau)^2}\\ \end{array} \right)\label{Md}\,. \end{eqnarray} The total stability matrix for the succession rotation, torsion, dissipation is given by $M=M_dM_tM_r$. \end{appendix}
\section{Introduction} We consider the system of two coupled nonlinear Schr\"odinger equations \begin{eqnarray} &&\imath {\mathcal U}_t+{\mathcal U}_{xx}+(\kappa {\mathcal U}{\mathcal U}^{\ast} +\chi {\mathcal V}{\mathcal V}^{\ast}){\mathcal U}=0, \nonumber \\ && \label{manakov} \\ &&\imath {\mathcal V}_t+{\mathcal V}_{xx} +( \chi{\mathcal U}{\mathcal U}^{\ast}+ \rho {\mathcal V}{\mathcal V}^{\ast}){\mathcal V}=0, \nonumber \end{eqnarray} where $\kappa,\chi,\rho$ are some constants. The integrability of this system is proven by Manakov \cite{ma74} only for the case $\kappa=\chi=\rho$, which we shall refer as {\it Manakov system}. The equations (\ref{manakov}) are important for a number of physical applications when $\chi$ is positive and all remaining constants equals to 1. For example for two-mode optical fibers $\chi=2$ \cite{ccp82} and for propagation of two modes in fibers with strong birefringence $\chi=\frac{2}{3}$ \cite{me87} and in the general case $\frac{2}{3}\leq \chi \leq 2$ for elliptical eigenmodes. The special value $\chi=1$ (Manakov system) corresponds to at least two possible cases, namelly the case of a purely electrostrictive nonlinearity or, in the elliptical birefringence, when angle between the major and minor axes of the birefingence ellipse is approximately $35^{o}$. The experimental observation of Manakov solitons in crystals is reported in \cite{ksaa96}. Recently Manakov model appear in the Kerr-type approximation of photorefractive crystals \cite{kpsv98}. The pulse-pulse collision between wavelength-division-multiplexed channels of optical fiber transmission systems are described with equations (\ref{manakov}) $\chi=2$, \cite{meg91, kmw96, hk95,ko97}. General quasi-periodic solutions in terms of $n$-phase theta functions for integrable Manakov system are derived in \cite{ahh93}, while a series of special solutions are given in \cite{allss95,puc98,phf98,pp99}. The authors of this paper discussed already quasi-periodic and periodic solutions associated to Lam\'e and Treibich-Verdier potentials for nonintegrable system of coupled nonlinear Schr\"odinger equations in frames of a special ansatz \cite{ceek95}. We also mention the method of constructing elliptic finite-gap solutions of the stationary KdV and AKNS hierarchy, based on a theorem due to Picard, is proposed in \cite{gw96,gw98b,gw98a} and the method developed by Smirnov in series of publications, the review paper\cite{sm94} and \cite{sm97,sm97a}. In the present paper we investigate integrable Manakov system being restricted to the system integrable in terms of ultraelliptic functions by introducing special ansats, which was recently apllied by {\it Porubov and Parker} \cite{pp99} to analyse special classes of elliptic solutions of the Manakov system $(\kappa=\chi=\rho=1)$. More precisely, we seek solution of (\ref{manakov}) in the form \begin{eqnarray} {\mathcal U}(x,t)=q_1(x) \,\mathrm{ exp}\left \{\imath a_1 t+\imath C_1\int\limits_{\cdot}^x \frac{{\mathrm d}x}{q_1^2(x)}\right\},\label{ansatz}\\ {\mathcal V}(x,t)=q_2(x) \,\mathrm{ exp}\left \{\imath a_2 t+\imath C_2\int\limits_{\cdot}^x \frac{{\mathrm d}x}{q_2^2(x)}\right\},\nonumber \end{eqnarray} where the functions $q_{1,2}(x)$ are supposed to be real and $a_1,a_2,C_1,C_2$ are real constants. Substituting (\ref{ansatz}) into (\ref{manakov}) we reduce the system to the equations \begin{eqnarray} \frac{\partial^2 q_1 }{\partial x^2} +\rho q_1^3+\chi q_1q_2^2-a_1q_1-\frac{C_1^2}{q_1^3}=0 \label{system2}\\ \frac{\partial^2 q_2 }{\partial x^2} +\kappa q_2^3+\chi q_2q_1^2-a_2q_2-\frac{C_2^2}{q_2^3}=0. \nonumber \end{eqnarray} The system (\ref{system2}) is the natural hamiltonian two-particle system with the hamiltonian of the form \begin{eqnarray} H&=& \frac12p_1^2+ \frac12p_2^2+\frac14( \rho q_1^4+2\chi q_1^2q_2^2+\kappa q_2^4) \cr&-&\frac12 a_1q_1^2 -\frac12a_2q_2^2+\frac12\frac{C_1^2}{q_1^2}+\frac12\frac{C_2^2}{q_2^2}, \end{eqnarray} where $p_1(t)= {\mathrm d}q_i(t)/dt$. These equations describe the motion of particles interacting with the quartic potential $Aq_1^4+Bq_1^2q_2^2+Cq_2^4$ and perturbed by inverse squared potential. Nowdays four nontrivial cases of complete integrabilty are known for nonperturbed potential (i) A:B:C= 1:2:1, (ii) A:B:C= 1:12:16, (iii) A:B:C= 1:6:1,(iv) A:B:C= 1:6:8. Cases (i), (ii) and (iii) are separable in respectively ellipsoidal, paraboloidal and Cartesian coordinates, while the case (iv) is separable in general sence \cite{rrg94}. The cases (ii) appears as one of the entries to polynomial hierarchy discussed in \cite{eekl93aa} the cases (iii) and (iv) are proved to be canonically equivalent under the action of Miura map restricted to the stationary of coupled KdV systems associated with fourth order Lax operator\cite{bef95}. Moreover all the cases (i)-(iv) permit the deformation of the potential by linear conbination of inverse squares and squares with certain limitations on the coefficients \cite{eekl93aa,bef95}. There are also known Lax representations for all these cases which yield hyperelliptic algebraic curves in the cases (i) and (ii) and 4-gonal curve in the cases (iii) and (iv). Although each from the system enumerated yield nontrivial classes of solutions of the system (\ref{manakov}) we shall discuss further only the case (i). The integrability of this case and separability in ellipsoidal coordinates was proved by {\it Wojciechowski} \cite{w85} (see also \cite{k89,t95}). We employ this result to integrate the system in terms of ultraelliptic functions (hyperelliptic functions of the genus two curve) and then execute reduction of hyperelliptic functions to elliptic ones by imposing additional constrains on the parameters of the system. The paper is organised as follows. In the first section we construct the Lax representation of the system, develop the genus two algebraic curve, which is associated to the system and reduce the problem to solution of the Jacobi inversion problem associated with genus two algebraic curve. In the section two develop the integration of the system in terms of {\it Kleinian hyperelliptic functions} which represent a natural generalization of Weierstrass elliptic functions to hyperelliptic curved of hidher genera; recently this realization of abelian functions was discussed in \cite{bel97b,bel97c,eel99}. We explain in the section the outline of the Kleinian realization of hyperelliptic functions and give the principle formulae for the case of genus two curve. In Section 4 we develop reduction of Kleinian hyperelliptic function to elliptic functions in terms of {\it Darboux coordinates} for the curve admiting additional involution. In this way a quasiperodic solution in terms of elliptic functions is obtained. In the last section we construct a set of elliptic periodic solutions on the basis of application of spectral theory for the Hill equation with elliptic potential. \section{Lax representation} The system $1:2:1$ $(\kappa=\chi=\rho=1)$ is a completely integrable hamiltonian system \begin{eqnarray} \frac{\partial^2 q_1 }{\partial x^2} +(q_1^2+q_2^2)q_1-a_1q_1-\frac{C_1^2}{q_1^3}=0,\cr \label{system}\\ \frac{\partial^2 q_2 }{\partial x^2} +(q_1^2+q_2^2)q_2-a_2q_2-\frac{C_2^2}{q_2^3}=0 \nonumber \end{eqnarray} with the Hamiltonian \begin{equation} H=\frac12\sum_{i=1}^2 p_i^2+\frac14\left(q_1^{2}+q_2^{2}\right)^2-\frac12a_1 q_1^2- \frac12 a_2 q_2^2+ \frac12\frac{C_1^2}{q_1^2}+ \frac12\frac{C_2^2}{q_2^2}, \label{H}\end{equation} where the variables $(q_1,p_1;q_{2},p_{2})$ are the canonicaly conjugated variables with respect to the standard Poisson bracket, $\{\cdot\;;\;\cdot\}$. This system permits the Lax representation (special case of Lax representation given in \cite{k98}). \begin{eqnarray} \frac{\partial L(\lambda)}{\partial \zeta}&=&[M(\lambda),L(\lambda)],\cr \quad L(\lambda)&=&\left( \begin{array}{cc} V(\lambda) & U(\lambda) \\ W(\lambda) & -V(\lambda) \end{array} \right),\quad M=\left( \begin{array}{cc} 0 & 1 \\ Q(\lambda) & 0 \end{array} \right) \label{lax} \end{eqnarray} is equvalent to the (\ref{system}), where $U(\lambda),W(\lambda),Q(\lambda)$ have the form \begin{eqnarray*} U(\lambda)&=&-a(\lambda)\left(1+\frac{1}{2}\frac{q_1^2}{\lambda-a_1} +\frac{1}{2}\frac{q_2^2}{\lambda-a_2}\right), \label{u} \\ V(\lambda)&=&-\frac12\frac{\mathrm{d}} {\mathrm{d\zeta}} U(\lambda) , \label{v} \\ W(\lambda)&=&a(\lambda)\left(-\lambda+\frac{q_1^2}{2}+\frac{q_2^2}{2} +\frac12\left(p_1^2+\frac{C_1^2}{ q_1^2}\right)\frac{1}{\lambda-a_1}\right.\cr&+&\left. \frac12\left(p_2^2+\frac{C_2^2}{q_2^2}\right) \frac{1}{ \lambda-a_2} \right) , \label{w} \\ Q(\lambda)&=&\lambda-q_1^2-q_2^2, \label{q} \end{eqnarray*} where $a(\lambda)=(\lambda-a_1)(\lambda -a_2)$. The Lax representation yields hyperelliptic curve $V=(\nu,\lambda)$, \[ \det(L(\lambda)-\frac12\nu 1_2)=0, \] where $1_2$ be $2\times2$ unit matrix and is given explicitly as \begin{eqnarray} \nu^2&=&4(\lambda-a_1)(\lambda-a_2)(\lambda^3-\lambda^2(a_1+a_2) +\lambda(a_1a_2-H)-F)\cr &-&C_1^2(\lambda-a_2)^2-C_2^2(\lambda-a_1)^2, \label{curve} \end{eqnarray} where $H$ is the hamiltonian (\ref{H}) and the second independent integral of motion $F$, $\{F;H\}=0$ is given as \begin{eqnarray} F&=&\frac14(p_1q_2-p_2q_1)^2+\frac12(q_1^2+q_2^2)(a_1a_2-\frac12a_2q_1^2 -\frac12a_1q_2^2)\cr &-&\frac12p_1^2a_2-\frac12p_2^2a_1 -\frac14\frac{(2a_2-q_2^2)C_1^2}{q_1^2} -\frac14\frac{(2a_1-q_1^2)C_2^2}{q_2^2}. \label{F}\end{eqnarray} We remark, that the parameters $C_i$ are linked with coordinates of the points $(a_i,\nu(a_i))$ by the formula \begin{equation} C_i^2=-\frac{\nu(a_i)^2}{(a_i-a_j)^2},\quad i,j=1,2.\label{ccc} \end{equation} Let us write the curve (\ref{curve}) in the form \begin{eqnarray} \nu^2=4\lambda^5+\alpha_4\lambda^4+\alpha_3\lambda^3+\alpha_2\lambda^2 +\alpha_1\lambda+\alpha_0,\label{curvecan} \end{eqnarray} where the {\it moduli} of the curve $\alpha_i$ are expressible in terms of physical parameters - level of energy $H$ and constants $a_1,a_2$, $C_1,C_2$ as follows \begin{eqnarray} \alpha_4&=&-8(a_1+a_2),\cr \alpha_3&=&-4H+4(a_1+a_2)^2+8a_1a_2,\cr \alpha_2&=&4H(a_1+a_2)-4F-C_1^2-C_2^2-8a_1a_2(a_1+a_2),\cr \alpha_1&=&4F(a_1+a_2)-4a_1a_2H+2C_1^2a_2+2C_2^2a_1 +4a_1^2a_2^2,\cr \alpha_0&=&-4a_1a_2F-C_1^2a_2^2-C_2^2a_1^2.\nonumber \end{eqnarray} Let us define new coordinates $\mu_1,\mu_2$ as zeros of the entry $ U(\lambda)$ to the Lax operator. Then \begin{eqnarray} q_1^2=2\frac{(a_1-\mu_1)(a_1-\mu_2)}{a_1-a_2},\quad q_2^2=2 \frac{ (a_2-\mu_1)(a_2-\mu_2)}{a_2-a_1}.\label{qcoord} \end{eqnarray} The definition of $\mu_1,\mu_2$ in the combination with the Lax representation comes to the equations \begin{equation} \nu_i=V(\mu_i)=-\frac12\frac{\partial}{\partial x}U(\mu_i),\quad i=1,2, \end{equation} wich can be transformed to the equations of the the form\footnote{In what follows we shall denote the integral bounds by the second coordinate of the curve $V=V(\nu,\lambda)$. (\ref{curve})} \begin{eqnarray} u_1=\int_{a_1}^{\mu_1}\mathrm{d}u_1 +\int_{a_2}^{\mu_2}du_1, \\ u_2=\int_{a_1}^{\mu_1}\mathrm{d}u_2 +\int_{a_2}^{\mu_2} du_2\label{jip} \end{eqnarray} where $\mathrm{d}u_{1,2}$ denote independent canonical holomorphic differentials \begin{equation} \mathrm{d}u_1= \frac{\mathrm{d}\lambda}{\nu},\quad \mathrm{d}u_2=\frac{\lambda \mathrm{d}\lambda}{\nu} .\label{hodbas} \end{equation} and $u_1=a,u_2=2x+b$ with the constants $a,b$ defining by the initial conditions. The integration of the problem is then reduces to the solving of the {\it Jacobi inversion problem } associated with the curve, which consist in the expession of the symmetric functions of $(\mu_1,\mu_2,\nu_1,\nu_2)$ as function of two complex variables $(u_1,u_2)$. \section[Kleinian hyperelliptic functions]{Exact solutions in terms of Kleinian hyperelliptic functions} In this section we give the trajectories of the system under consideration in terms of Kleinan hyperelliptic functions (see, e.g. \cite{ba97,bel97c}), being associated with the algebraic curve of genus two (\ref{curvecan}) which can be also written in the form \begin{eqnarray} \nu^2&=4\prod_{i=0}^{4}(\lambda-\lambda_{i}), \label{gen2} \end{eqnarray} where $\lambda_{i}\neq \lambda_{i}$ are branching points. At all real branching points the closed intervals $[\lambda_{2i-1},\lambda_{2i}],i=0,\ldots 4$ will be referred further as lacunae \cite{zmnp80,mm75}. Let us equip the curve with a homology basis $({\mathfrak a}_1,{\mathfrak a}_2; {\mathfrak b}_1, {\mathfrak b}_2)\in H_1(V,{\mathbb Z})$ and fix the basis in the space of holomorphic differentials as in (\ref{hodbas}). The associated canonical meromorphic differentials of the second kind $\mathrm{ d}\boldsymbol {r}^T=(\mathrm{ d}r_1,{\mathrm d} r_2)$ have the form \begin{equation}{\mathrm d}r_1=\frac{\alpha_3\lambda+2\alpha_4\lambda^2+12\lambda^3}{ 4\nu}d\lambda,\qquad {\mathrm d}r_2=\frac{\lambda^2}{ \nu}d\lambda.\label{rr} \end{equation} The $2\times 2$ matrices of their periods, \begin{eqnarray*} 2\omega&=&\left(\oint_{{\mathfrak a}_k}{\mathrm d} u_l\right)_{k,l=1,2},\quad 2\omega'=\left(\oint_{{\mathfrak b}_k}{\mathrm d} u_l\right)_{k,l=1,2},\\ 2\eta&=&\left(\oint_{{\mathfrak a}_k}{\mathrm d} r_l\right)_{k,l=1,2},\quad 2\eta'=\left(\oint_{{\mathfrak b}_k}{\mathrm d} r_l\right)_{k,l=1,2} \end{eqnarray*} satisfy the equations, \[\omega'\omega^T-\omega{\omega'}^T=0,\quad \eta'\omega^T-\eta{\omega'}^T=-\frac{\imath\pi}{2}1_2,\quad \eta'\eta^T-\eta{\eta'}^T=0, \] which generalizes the Legendre relations between complete elliptic integrals to the case $g=2$. The fundamental $\sigma$ function in this case is a natural generalization of the Weierstrass elliptic $\sigma$ function and is defined as follows \begin{eqnarray*} \sigma(\boldsymbol{u})&=&\frac{\pi}{\sqrt{\mathrm{det}(2\omega)}} \frac{\epsilon}{\sqrt[4]{\prod_{1\leq i<j\leq 5}(a_i-a_j)}}\\ &\times&\exp\left\{\boldsymbol{ u}^T\eta(2\omega)^{-1}\boldsymbol{u}\right\} \theta[\varepsilon]((2\omega)^{-1} \boldsymbol{ u}|\omega'\omega^{-1}), \end{eqnarray*} where $\epsilon^8=1$ and $\theta[\varepsilon](\boldsymbol{ v}|\tau)$ is the $\theta$ function with an odd characteristic $[\varepsilon]=\left[\begin{array}{cc}\varepsilon_1&\varepsilon_2\\ \varepsilon_1'&\varepsilon_2'\end{array}\right]$, which is the characteristic of the vector of Riemann constants, \[\theta[\varepsilon](\boldsymbol{ v}|\tau)=\sum_{\boldsymbol{ m}\in{\mathbb Z}^2}\mathrm{\exp}\;\imath\pi\left\{ (\boldsymbol{ m}+{\boldsymbol\varepsilon})^T\tau (\boldsymbol{ m}+{\boldsymbol \varepsilon})+2 (\boldsymbol{ v}+{\boldsymbol \varepsilon}')^T\tau (\boldsymbol{ m}+{\boldsymbol\varepsilon})\right\}. \] Alternatively the $\sigma $ function can be defined by its expansion near $\boldsymbol{u}=0$ \begin{equation} \sigma (\boldsymbol{u})=u_1+\frac{1}{24} \alpha_2u_1^3-\frac{1}{3}u_2^3+o(\boldsymbol{u}^5 )\label{ex} \end{equation} and the further terms can be computed with the help of bilinear differential equation \cite{ba07}. The $\sigma$-function posses the following periodicity property: put \[ \boldsymbol{E}(\boldsymbol{ m},\boldsymbol{ m}')=\eta\boldsymbol{ m} +\eta'\boldsymbol{ m}',\quad\text{and}\quad \boldsymbol{\Omega}(\boldsymbol{ m},\boldsymbol{ m }') =\omega\boldsymbol{ m} +\omega' \boldsymbol{ m}', \] where $\boldsymbol{ m},\boldsymbol{ m}'\in \mathbb{Z}^{n}$, then \begin{align*} &\sigma[\varepsilon](\boldsymbol{z}+ 2{\boldsymbol\Omega} (\boldsymbol{ m }, \boldsymbol{ m }'),\omega,\omega') =\mathrm{exp} \big\{ 2\boldsymbol{E}^T (\boldsymbol{ m},\boldsymbol{ m}') \big({\boldsymbol z}+ \boldsymbol{\Omega}(\boldsymbol{ m }, \boldsymbol{ m }')\big)\big\}\\ &\times \mathrm{exp} \{ -\pi \imath {\boldsymbol m }^T{\boldsymbol m}' -2\pi \imath {\boldsymbol\varepsilon }^T{\boldsymbol m}' \} \sigma[\varepsilon]( {\boldsymbol z},\omega,\omega') \end{align*} As modular function the Kleinian $\sigma$-function is invariant under the transformation of the symplectic group, what represents the important characteristic feature. We introduce the Kleinian hyperelliptic functions as second logarithmic derivatives \begin{eqnarray*} \wp_{11}(\boldsymbol{ u})&=&-\frac{\partial^2}{ \partial u_1^2}\mathrm{ ln}\; \sigma(\boldsymbol{ u}),\quad \wp_{12}(\boldsymbol{ u})=-\frac{\partial^2}{\partial u_1\partial u_2}\mathrm{ ln}\; \sigma(\boldsymbol{ u}),\cr \wp_{22}(\boldsymbol{ u})&=&-\frac{\partial^2}{\partial u_2^2} \mathrm{ ln}\; \sigma(\boldsymbol{ u}). \end{eqnarray*} The multi-index symbols $\wp_{i,j,k}$ etc. are defined as logarithmic derivatives by the variable $u_i,u_j,u_k$ on the corresponding indices $i,j,k$ etc. The principal result of the theory is the formula of Klein, which reads in the case of genus two as follows \begin{eqnarray} &&\sum_{k,l=1}^2\wp_{kl}\left(\int_{\infty}^{\mu} {\mathrm d}{\mathbf u}- \int_{\infty}^{\mu_1 } {\mathrm d}{\mathbf u}- \int_{\infty}^{\mu_2 } {\mathrm d}{\mathbf u}\right)\mu^{k-1}\mu_i^{l-1}\cr &=&\frac{F(\mu,\mu_i)+2\nu\nu_i}{4(\mu-\mu_i)^2}, \quad i=1,2,\label{klein} \end{eqnarray} where \begin{equation} F(\mu_1,\mu_2)=\sum_{r=0}^2\mu_1^r\mu_2^r[2\alpha_{2r} +\alpha_{2r+1}(\mu_1+\mu_2)].\label{fx1x2} \end{equation} By expanding these equalities in the vicinity of the infinity we obtain the complete set of the relations for hyperelliptic functions. The first group of the relations represents the solution of the Jacobi inversion problem in the form \begin{equation} \lambda^2-\wp_{22}(\boldsymbol{u})\lambda -\wp_{12}(\boldsymbol{u})=0,\label{bolza2}\end{equation} that is, the pair $(\mu_1,\mu_2)$ is the pair of roots of (\ref{bolza2}). So we have \begin{equation}\wp_{22}(\boldsymbol{u}) =\mu_1+\mu_2,\;\wp_{12}(\boldsymbol{u})=-\mu_1\mu_2.\label{b1}\end{equation} The corresponding $\nu_i$ is expressed as \index{$\wp$ function!fundamental relation!for genus two} \begin{equation} \nu_i=\wp_{222}(\boldsymbol{u})\mu_i+\wp_{122}(\boldsymbol{ u}),\quad i=1,2. \label{y2} \end{equation} The functions $\wp_{22},\wp_{12}$ are called basis functions. The function $\wp_{11}(\boldsymbol{u})$ can be also espressed as symmetric function of $\mu_1,\mu_2$ and $\nu_1,\nu_2$: \begin{equation} \wp_{11}(\boldsymbol{u})= \frac{F(\mu_1,\mu_2)-2\nu_1\nu_2}{4(\mu_1-\mu_2)^2}, \label{b2} \end{equation} where $F(\mu_1,\mu_2)$ is given in (\ref{fx1x2}). Further from (\ref{y2}) we have \begin{eqnarray} \wp_{222}(\boldsymbol{ u})&=&\frac{\nu_1-\nu_2}{ \mu_1-\mu_2},\quad \wp_{221}(\boldsymbol{ u})=\frac{\mu_1\nu_2-\mu_2\nu_1}{ \mu_1-\mu_2}, \nonumber \\ \wp_{211}(\boldsymbol{ u})&=&-\frac{\mu_1^2\nu_2-\mu_2^2\nu_1}{ \mu_1-\mu_2}, \nonumber \\ \wp_{111}(\boldsymbol{ u})&=&\frac{\nu_2\psi(\mu_1,\mu_2)-\nu_1 \psi(\mu_2,\mu_1)}{ 4(\mu_1-\mu_2)^3}, \label{thirdder} \end{eqnarray} where \begin{eqnarray} \psi(\mu_1,\mu_2) &=& 4\alpha_0 + \alpha_1(3\mu_1 + \mu_2) + 2\alpha_2\mu_1(\mu_1 +\mu_2) \nonumber \\ &+& \alpha_3\mu_1^2(\mu_1 +3\mu_2) + 4\alpha_4\mu_1^2\mu_2 +4\mu_1^2\mu_2(3\mu_1 +\mu_2). \nonumber \end{eqnarray} The next group of the relations, which can be drived by the expanding of the equations (\ref{klein}) are the pairwise products of the $\wp_{ijk}$ functions being expressed in terms of $\wp_{22},\wp_{12},\wp_{11}$ and constants $\alpha_s$ of the defining equation (\ref{gen2}). We give here only basis equations \begin{eqnarray*} \wp_{222}^2&=4\wp_{22}^3+4\wp_{12}\wp_{22}+\alpha_4\wp_{22}^2+4\wp_{11} +\alpha_3\wp_{22}+\alpha_2,\\ \wp_{222}\wp_{122}&=4\wp_{12}\wp_{22}^2 +2\wp_{12}^2-2\wp_{11}\wp_{22}+\alpha_4\wp_{12}\wp_{22} \\ &+\frac12\alpha_3\wp_{12}+\frac12\alpha_1 , \\ \wp_{122}^2&=4\wp_{22}\wp_{12}^2-4\wp_{11}\wp_{12}+ \alpha_4\wp_{12}^2-\alpha_0. \end{eqnarray*} All such expressions may be rewritten in the form of an {\it extended cubic relation} as follows. For arbitrary $\boldsymbol{ l},\boldsymbol { k}\in {\mathbb C}^4$ the following formula is valid \cite{ba07} \begin{equation}\boldsymbol{l}^T\pi\pi^T \boldsymbol{k}=-\frac14{\det}\;\left(\begin{array}{cc}H&\boldsymbol{l}\\ \boldsymbol{k}^T&0\end{array}\right),\label{kum1}\end{equation} where $\pi^T=(\wp_{222},-\wp_{221},\wp_{211},-\wp_{111})$ and $H$ is the $4\times4$ matrix: \begin{equation} H= \left ( \begin{array}{cccc} \alpha_0& \frac{1}{2} \alpha_1&-2 \wp_{11}&-2 \wp_{12} \\ \frac{1}{2} \alpha_1& \alpha_2+4 \wp_{11}& \frac{1}{2} \alpha_3+ 2 \wp_{12}&-2 \wp_{22} \\-2 \wp_{11}& \frac{1}{2} \alpha_3+2 \wp_{12}& \alpha_4+4 \wp_{22}&2 \\ -2 \wp_{12}&-2 \wp_{22}&2&0 \end{array} \right) . \end{equation} The vector $\pi$ satisfies the equation $H\pi=0$, and so the functions $\wp_{22},\wp_{12}$ and $\wp_{11}$ are related by the equation \begin{equation}\mathrm{ det}\; H=0.\label{kum5} \end{equation} The equation (\ref{kum5}) defines the quartic Kummer surface $\mathbb K$ in $\mathbb C^3$ \cite{hu05}. The next group of the equations, which is derived as the result of expansion of the equalities (\ref{klein}) are the expressions of four index symbols $\wp_{ijkl}$ as quadrics in $\wp_{ij}$ \begin{eqnarray} &\wp_{2222}=6\wp_{22}^2+\frac12\alpha_3+\alpha_4\wp_{22} +4\wp_{12},\label{eeq1}\\ &\wp_{2221}=6\wp_{22}\wp_{12}+\alpha_4\wp_{12}-2\wp_{11},\label{eeq3}\\ &\wp_{2211}=2\wp_{22}\wp_{11}+4\wp_{12}^2+ \frac12\alpha_3\wp_{12}.\label{eeq5}\\ &\wp_{2111}=6\wp_{12}\wp_{11}+\alpha_2\wp_{12} -\frac12\alpha_1\wp_{22}-\alpha_0,\label{eeq4}\\ &\wp_{1111}=6\wp_{11}^2-3\alpha_0\wp_{22} +\alpha_1\wp_{12}+\alpha_2\wp_{11}- \frac12\alpha_0\alpha_4+\frac18\alpha_1\alpha_3.\label{eeq2} \end{eqnarray} These equations can be identified with completely integrable partial differential equations and dynamical systems, which are solved in terms of Abelian functions of hyperelliptic curve of genus two. In particular, the first two equations represent the KdV hierarhy with ``times" $(t_1,t_2)=(u_2,u_1)=(x,t)$, \begin{equation}{\mathcal X}_{k+1}[{\mathsf U}]={\mathcal R}{\mathcal X}_{k}[{\mathsf U}] \end{equation} where ${\mathcal R}=\partial_x^2- {\mathsf U}+c -\frac12{\mathsf U}_x\partial^{-1}$, $c=\alpha_4/12$ is the Lenard recursion operator. The first two equations from the hierarchy are \begin{equation} {\mathsf U}_{t_1}={\mathsf U}_{x},\quad {\mathsf U}_{t_2}=\frac12({\mathsf U}_{xxx}-6{\mathsf U}_{x}{\mathsf U}), \label{kdv} \end{equation} the second equation is the KdV equation, which is obtained from (\ref{eeq1}) as the result of differentiation by $x=u_2$ and setting ${\mathsf U}=2\wp_{22}+\alpha_4/6$. The equation (\ref{eeq1}) plays role of the stationary equation in the hierarchy and is obtained as the result of action of the recursion operator. Let us introduce finaly the {\it Baker-Akhiezer} function, which in the frames of the formalizm developed is expressible in terms of the Kleinian $\sigma$-function as follows \begin{equation} \Psi(\lambda,\boldsymbol{u})= \frac{ \sigma\left(\int_{\infty}^{\lambda}{\mathrm d} \boldsymbol{ u}- {\mathbf u} \right)} {\sigma(\boldsymbol{u}) } \mathrm {exp}\left\{ \int_{\infty}^{\lambda}{\mathrm d} {\mathbf r}^T \boldsymbol{u} \right\},\label{BAF} \end{equation} where $\lambda$ is arbitrary and $\boldsymbol u$ is the Abel image of arbitrary point $(\nu_1,\mu_1)\times (\nu_2,\mu_2)\in V \times V $. It is straighforward to show by the direct calculation, being bases on the usage of the relations for three and four-index Kleinian $\wp$--functions that $\Psi(\lambda,\boldsymbol{u})$ satisfy to the Schr\"odinger equation \begin{equation} (\frac{\partial^2}{{\partial u_2}^2}-2\wp_{22}(\boldsymbol{u})) \Psi(\lambda,\boldsymbol{u})= \left(\lambda+\frac14\alpha_{4}\right)\Psi(\lambda,\boldsymbol{u})\label{sch} \end{equation} for all $(\nu,\mu)$. Now we are in position to write the solution of the of the system in terms of Kleinian $\sigma$-functions and identify the constants in terms of the moduli of the curve. Using (\ref{b1}),(\ref{qcoord}) the solutions of (\ref{system}) have the following form in terms of Kleinian functions $\wp_{22}(\boldsymbol{ u}), \wp_{12}(\boldsymbol{ u})$ \begin{eqnarray} q_1^2&=&2\frac{a_1^{2}-\wp_{22}(\boldsymbol u)a_1-\wp_{12}(\boldsymbol u)}{a_1-a_2}, \cr q_2^2&=&2\frac{a_2^{2}-\wp_{22}(\boldsymbol u)a_2-\wp_{12}(\boldsymbol u)}{a_2-a_1}, \label{solution} \end{eqnarray} where the vector $\boldsymbol {u}^T=(a,2x+b)$. \section[Quasi-periodic elliptic solutions] {Periodic solutions expressed in terms of elliptic functions of different moduli} We consider in this section the reduction Jacobi (see e.q.\cite{kr03} ) of hyperelliptic integrals to elliptic ones, when the hyperelliptic curve $V$ has the form \begin{equation} w^2=z(z -1)(z -\alpha )(z -\beta )( z -\alpha \beta ) \label{curver} \end{equation} The curve (\ref{curver}) covers two-sheetedly two tori $$\pi _{\pm }:V=(w,z)\rightarrow E_{\pm }=(\eta _{\pm },\xi _{\pm }),$$ \begin{equation} \eta _{\pm }^2=\xi _{\pm }(1-\xi _{\pm })(1-k_{\pm }^2\xi _{\pm }) \end{equation} with Jacobi moduli \begin{equation} k_{\pm }^2=-\frac{(\sqrt{\alpha }\mp \sqrt{\beta })^2}{(1-\alpha )(1-\beta )} , \label{jacmod} \end{equation} The covers $\pi _{\pm }$ are described by the formulae \begin{eqnarray} \eta _{\pm }=-\sqrt{(1-\alpha )(1-\beta )}\frac{z\mp \sqrt{\alpha \beta }}{ (z-\alpha )^2(z-\beta )^2}w, \label{r1} \\ \xi=\xi _{\pm}=\frac{(1-\alpha )(1-\beta )z}{(z-\alpha )(z-\beta )}. \label{r2} \end{eqnarray} The following formula is valid for the reduction of holomorphic hyperelliptic differential to the elliptic ones: \begin{equation} \frac{d\xi _{\pm }}{\eta _{\pm }}=-\sqrt{(1-\alpha )(1-\beta )}(z\mp \sqrt{ \alpha \beta })\frac{\mathrm{d} z}{w}. \label{r3} \end{equation} Suppose that the spectral curve (\ref{curvecan}) admits the symmetry of the (\ref{curver}) and apply the discussed reduction case to the problem. Then the equations of the Jacobi inversion problem (\ref{jip}) can be rewritten in the form \begin{eqnarray} \sqrt{(1-\beta)(1-\alpha)}\sum_{i=1}^2\int_{z_0}^{z_i} (z -\sqrt{\alpha\beta})\frac{\mathrm{d}z} {w}=2u_{+}, \label{j11} \\ \sqrt{(1-\beta)(1-\alpha)}\sum_{i=1}^2\int_{x_0}^{z_i} (z +\sqrt{\alpha\beta})\frac{\mathrm{d}z} {w}=2u_{-} .\label{j22} \end{eqnarray} with $(\nu_i,\mu_i)=(2w_i,z_i)$ and \begin{equation} u_{\pm }=-\sqrt{(1-\alpha )(1-\beta )}(u_2\mp \sqrt{\alpha \beta }u_1) \end{equation} Reduce in (\ref{j11},\ref{j22}) hyperelliptic integrals to elliptic ones according to (\ref{r1},\ref{r2}). \begin{eqnarray*} \int_{0}^{\sqrt{\xi(\mu_1)}}\frac{\mathrm{d}x}{\sqrt{(1-x^2)(1-k^2_{\pm}x^2)}} +\int_{0}^{\sqrt{\xi(\mu_2)}}\frac{\mathrm{d}x} {\sqrt{(1-x^2)(1-k^2_{\pm}x^2)}} =u_{\pm}, \end{eqnarray*} One can further express the symmetric functions of $\mu_1,\mu_2, \nu_1, \nu_2$ on $V\times V$ in term of elliptic functions of tori $E_{\pm}$. To this end we introduce the {\it Darboux coordinates} (see \cite{hu05}, p.105 ) \begin{eqnarray} X_1=\mbox{sn}(u_{+},k_{+})\mbox{sn}(u_{-},k_{-}),\cr X_2=\mbox{cn}(u_{+},k_{+})\mbox{cn}(u_{-},k_{-}), \label{darb} \\ X_3=\mbox{dn}(u_{+},k_{+})\mbox{dn}(u_{-},k_{-}), \nonumber \end{eqnarray} where $\mbox{sn}(u_{\pm },k_{\pm }),\mbox{cn}(u_{\pm },k_{\pm }),\mbox{dn }(u_{\pm },k_{\pm })$ are standard Jacobi elliptic functions. We apply further the addition theorem for Jacobi elliptic functions, \begin{eqnarray} \mbox{sn}(u_1+u_2,k)=\frac{s_1^2-s_2^2}{s_1c_2d_2-s_2c_1d_1},\cr \mbox{cn }(u_1+u_2,k)=\frac{s_1c_1d_2-s_2c_2d_1}{s_1c_2d_2-s_2c_1d_1},\cr \mbox{dn} (u_1+u_2,k)=\frac{s_1d_1c_2-s_2d_2s_1}{s_1c_2d_2-s_2c_1d_1}, \nonumber \end{eqnarray} where we denoted $s_i=\mbox{sn}(u_i,k),c_i=\mbox{cn}(u_i,k),d_i=\mbox{ dn}(u_i,k) $, $i=1,2$ and formulae (\ref{b1},\ref{b2}) for the Kleinian hyperelliptic functions. The straightforward calculations lead to the formulae \begin{eqnarray} X_1=-\frac{(1-\alpha )(1-\beta )(\alpha \beta +\wp _{12})}{(\alpha +\beta )(\wp _{12}-\alpha \beta )+\alpha \beta \wp _{22}+\wp _{11}}, \label{x1} \cr X_2=-\frac{(1+\alpha \beta )(\alpha \beta -\wp _{12})-\alpha \beta \wp _{22}-\wp _{11}}{(\alpha +\beta )(\wp _{12}-\alpha \beta )+\alpha \beta \wp _{22}+\wp _{11}}, \label{x2} \\ X_3=\frac{\alpha \beta \wp _{22}-\wp _{11}}{(\alpha +\beta )(\wp _{12}-\alpha \beta )+\alpha \beta \wp _{22}+\wp _{11}}. \label{x3}\nonumber \end{eqnarray} The formulae (\ref{x2}) can be inverted as follows \begin{eqnarray} \wp_{11}=(B-1)\frac{A(X_2+X_3)-B(X_3+1)}{X_1+X_2-1}, \label{wp11} \\ \wp_{12}=(B-1)\frac{1+X_1-X_2}{X_1+X_2-1}, \label{wp12} \\ \wp_{22}=\frac{A(X_2-X_3)+B(X_3-1)}{X_1+X_2-1}, \label{wp22} \end{eqnarray} where $A=\alpha +\beta $, $B=1+\alpha \beta $. The obtained results permit to present few solutions in elliptic functions of the initial problem, which are quasi-periodic in $\zeta$. Using (\ref{wp12}) and (\ref{wp22}) for solutions of the (\ref{system}) in the form (\ref{solution}) we have \begin{eqnarray*} &&q_1^2=2\frac{1}{a_1-a_2}\left(a_1^{2}- \frac{A(X_2-X_3)+B(X_3-1)}{X_1+X_2-1}a_1\right.\cr&&\left.\qquad- (B-1)\frac{1+X_1-X_2}{X_1+X_2-1}\right), \\ &&q_2^2=2\frac{1}{a_2-a_1}\left(a_2^{2}- \frac{A(X_2-X_3)+B(X_3-1)}{X_1+X_2-1}a_2\right.\cr&&\left.\qquad- (B-1)\frac{1+X_1-X_2}{X_1+X_2-1}\right),\end{eqnarray*} where \begin{equation} u_{\pm }=-2\sqrt{(1-\alpha )(1-\beta )}(x\mp c ) \end{equation} and $c$ is the constant depending on initial conditions. We also remark, that the derived quasi perodic solution was associated with the Jacobi reduction case in which the ultraelliptic integrals were reduced to elliptic ones by the aid of second order substitution. This means on the language of two-dimensional $\theta$-functions, that the associated period matrix is equvalent to the matrix with the off-diagonal element $\tau_{12}=\frac12$. Such the reduction case was considered in various places (see e.g.\cite{bbeim94}). Solutions of this type for nonlinear Schr\"odinger equation ($\sigma=0$) are recently obtained in \cite{ch95}. The anologous technique can be carried out for other well documented case of reduction , when $\tau_{12}=1/N$ and the $N=3,4,\ldots$. In general such the reduction can be caried out for covers of arbitrary degree within the Weierstrass-Poincar\'e reduction theory (see e.g. \cite{kr03,bbeim94}). \section{Elliptic periodic solutions} In this section we develop a method (see also \cite{k89,ek94,ee94b}) which allows us to construct periodic solutions of (\ref{system}) in a straightforward way based on the application of spectral theory for the Schr\"odinger equation with elliptic potentials \cite{amm77,mm75}. We start with the formula (\ref{eeq1}) and with equation with equation for Baker function $\Psi(\lambda;\boldsymbol{ u})$. \begin{eqnarray} &&\frac {\mathrm{d}^2} {\mathrm{d} x^2} \Psi(\lambda,\boldsymbol{ u}) -{\mathsf U} \Psi(x,\boldsymbol{ u}) = (\lambda+\frac{\alpha_{4}}{4})\Psi(\lambda,\boldsymbol{ u}), \label{baker} \end{eqnarray} where we identify the potential $${\mathsf U}=2\wp_{22}+\frac16\alpha_{4}.$$ We assume, without loosing generality, that the associated curve has the property $\alpha_4=0$. To make this assumption applicable to the initial curve of the system (\ref{system}) being derived from the Lax representation, we undertake the shift of the spectral parameter, \begin{equation} \lambda\longrightarrow \lambda+\Delta,\qquad \Delta=\frac25a_1+\frac25a_2. \label{shift} \end{equation} Suppose, that ${\mathsf U}$ be two gap Lam\'e or two gap {\it Treibich-Verdier} potential, what means, that \begin{equation} {\mathsf U}(x)=2\sum_{i=1}^N \wp(x-x_i) \label{TVP}, \end{equation} where $\wp(x)$ is standard Weierstrass elliptic functions with periods $2\omega,2\omega'$ and numbers $x_i$ takes the values from the set $\{ 0,\omega_1=\omega,\omega_2=\omega+\omega',\omega_3=\omega'\}$. It is known, that the set of such the potentials is exhausted by six potentials\cite{tv90,ee95a} \begin{eqnarray} {\mathsf U}_3(x)&=&6\wp(x), \label{L3} \\ {\mathsf U}_4(x)&=&6\wp(x)+2\wp(x+\omega_i),\quad i=1,2,3,\label{TV4}\\ {\mathsf U}_5(x)&=&6\wp(x)+2\wp(x+\omega_i)+2\wp(x+\omega_j), \cr&& \qquad\qquad i\neq j=1,2,3, \label{TV5}\\ {\mathsf U}_6(x)&=&6\wp(x)+6\wp(x+\omega_i),\quad i=1,2,3,\nonumber\\ {\mathsf U}_8(x)&=&6\wp(x)+2\sum_{i=1}^3\wp(x+\omega_i), \nonumber\\ {\mathsf U}_{12}(x)&=&6\wp(x)+6\sum_{i=1}^3\wp(x+\omega_i), \nonumber \end{eqnarray} where the subscript shows the number of $2\wp$ functions involved and display the degree of the cover of the associated genus two curve over elliptic curve. Because the last three potentials can be obtained from the first three by Gauss transform we shall call the first three as {\it basis potentials}. The potential (\ref{L3}) is two gap Lam\'e potential, which is associated with three sheeted cover of elliptic curve; the potentials (\ref{TV4},\ref{TV5}) are Treibich-Verdier potentials \cite{ve90,tv90} associated with four and five sheeted cover correspondingly. To display the class of periodic solutions of system (\ref{system}) we introduce the {\it generalized Hermite polynomial} ${\mathcal F}(x,\lambda)$ by the formula \begin{equation} {\mathcal F}(x,\lambda)=\lambda^2-\pi_{22}(x)\lambda-\pi_{12}(x) \end{equation} with $\pi_{22}(x)$ and $\pi_{12}(x)$ given as follows \begin{eqnarray} \pi_{22}(x) &=&\sum_{j=1}^N \wp (x - x_j) + \frac13 \sum_{j=1}^5 \lambda_j,\label{tr1} \cr \pi_{12}(x) &=&-3\,\sum_{i<j} \wp(x - x_i) \wp(x -x_j) - \frac{Ng_2}{8} \nonumber\\ &&-\frac16 \sum_{i<j} \lambda_i \lambda_j + \frac{1}{ 6}\left(\sum_{j=1}^5 \lambda_j^2 \right) \label{tr2} \end{eqnarray} where $x_i$ are half-periods and $N$ is the degree of the cover (see for example \cite{ek94}). The introduction of this formula is based on the possibility to compute the symmetric function $\mu_1\mu_2$ in terms of differential polynomial of the first one with the help of the equation (\ref{eeq1}), which serves in this context as the ``trace formula" \cite{zmnp80}. The solutions of the system (\ref{system}) are then given as \begin{equation} q_1^2(x)=2\frac{{\mathcal F}(x,a_{1}-\Delta)} {a_{1}-a_{2}} , \quad q_2^2(x)=2\frac{{\mathcal F}(x,a_{2}-\Delta)} {a_{2}-a_{1}} ,\label{answer} \end{equation} The final formula for the solutions of the system (\ref{manakov}) then reads \begin{eqnarray} {\mathcal U}(x,t)=\,\sqrt{2\frac{{\mathcal F}(x,a_{1}-\Delta)} {a_{1}-a_{2}}} \mathrm{ exp}\left \{\imath a_1 t-\frac12\nu(a_i-\Delta)\int\limits_{\cdot}^x \frac{{\mathrm d}x}{{\mathcal F}(x,a_1-\Delta)}\right\},\cr \label{final}\\ {\mathcal V}(x,t)=\sqrt{2\frac{{\mathcal F}(x,a_{2}-\Delta)} {a_{2}-a_{1}}} \mathrm{ exp}\left \{\imath a_2 t-\frac12\nu(a_2-\Delta)\int\limits_{\cdot}^x \frac{{\mathrm d}x}{{\mathcal F}(x,a_2-\Delta)}\right\} ,\nonumber \end{eqnarray} where we used (\ref{answer}) and (\ref{ccc}). It is important to remark for our consideration, that if the potential is known, then the associated algebraic curve of genus two can be described with the help of the Novikov equation \cite{no74}. Let us consider the two-gap potential for normalized by its expansion near the singular point as \begin{equation} {\mathsf U}(x) = \frac{6}{ x^2} + a x^2 + b x^4 + c x^6 + d x^8 + O(x^{10}). \label{decomposition} \end{equation} Then the algebraic curve associated with this potential has the form \cite{be89b} \begin{eqnarray} \nu^2&=&\lambda^5 - \frac{5\cdot7}{2} a\lambda^3 +\frac{3^2\cdot7}{2} b\lambda^2 \cr &+& \left( \frac{3^4\cdot7}{8} a^2 +\frac{3^3\cdot11}{4}c \right)\lambda -\frac{3^4\cdot17}{4}ab+\frac{3^2\cdot11\cdot13}{2}d\label{curvelame}. \end{eqnarray} We shall consider below examples of genus two curves, which are associated with the two gap elliptic potentials (\ref{L3}), (\ref{TV4}) and (\ref{TV5}). Consider the potential ${\mathsf U}_3$ and construct the associated curve (\ref{curvelame}) \begin{equation}{\mathcal L}^2=(\lambda^2-3g_2)(\lambda+3e_1) (\lambda+3e_2)(\lambda+3e_3),\label{curve3}\end{equation} The Hermite polynomiaal ${\mathcal F}_3(\wp(x),\lambda)$ \cite{ww86} associated to the Lame potential (\ref{L3}), which is already normalized as in (\ref{decomposition}) has the form \begin{equation} {\mathcal F}_3(\wp(x),\lambda)=\lambda^{2}- 3\wp(x)\lambda + 9\wp^{2}(x)-\frac{9}{4}g_{2}. \label{HerPol} \end{equation} Then the finite and real solution to the system (\ref{system}) is given by the formula (\ref{answer}) with the Hermite polynomial depending in the argument $x+\omega'$ (the shift in $\omega'$ provides the holomorphity of the solution). The solution is real under the choice of the arbitrary constants $a_{1,2}$ in such way, that the constants $a_{1,2}-\Delta$ lie in {\it different } lacunae. According to (\ref{ccc}) the constants $C_{i}$ are then given as \[C_i^2=-\frac{4\nu^2(a_i-\Delta)}{(a_i-a_j)^2}\label{cccc}, \] where $\Delta$ is the shift (\ref{shift}) and $\nu$ is the coordinate of the curve (\ref{curve3}) and the integrals $H$ and $F$ have the following form \begin{eqnarray} && H=\frac{1}{25}\left(a_1+a_2\right)^3+\frac{21}{4} g_2, \nonumber \\ && F=\frac{1}{25}\left(a_{{1}}+a_{{2}}\right)^{3}-\frac{1}{4}C_{1}^{2} -\frac{1}{4}C_{2}^{2}-\frac{27}{4}g_{{3}}-\frac{21}{20}g_{{2}} \left(a_{1}+a_{2}\right). \nonumber \end{eqnarray} These results are in complete agreement with solutions obtained in \cite{pp99} by introducing the ansatz of the form $$q_i(x)=\sqrt{A_i\wp(x)^2+B_i\wp(x)+C_i},\quad i=1,2$$ with the constants $A_i,B_i,C_i$ which are defined from the compatibility condition of the ansatz with the equations of motion. In the foregoing examples we are considering the solutions of the form $$q_i(x)=\sqrt{{\mathcal Q}_i(\wp(x))}, $$ where ${\mathcal Q}_i$ are rational functions of $\wp(x)$. Consider with this purpose the Treibich Verdier potential \begin{equation} {\mathsf U}_4(x)=6\wp(x)+2\wp(x+\omega_1)-2e_1,\label{TV4} \end{equation} associated with four sheeted cover. The potential is normalized according to (\ref{decomposition}). The associated spectral curve is of the form \begin{eqnarray}{\nu}^2&=&4(\lambda+6e_1)\prod_{k=1}^4 (\lambda-\lambda_k) \label{ctv4}\\ \lambda_{1,2}&=&e_3+2e_2\pm 2\sqrt{(5e_3+7e_2)(2e_3+e_2)}\\ \lambda_{3,4}&=&e_2+2e_3\pm 2\sqrt{(5e_2+7e_3)(2e_2+e_3)}\nonumber, \end{eqnarray} The Hermite polynomials are given by the formula \begin{eqnarray} {\mathcal F}(x,\lambda)&=& \lambda^2-(3\wp(x)+\wp(x+\omega_1)-e_1)\lambda\label{hertv4}\\&+& 9\wp(x)(\wp(x)+\wp(x+\omega)-e_1)-3e_1\wp(x+\omega_1)\cr&+& \frac{9}{4}g_2-51e_1^2; \nonumber \end{eqnarray} The finite real solution of (\ref{system}) results the substitution this Hermite polynomial $ {\mathcal F}(x+\omega',\lambda)$ into (\ref{answer}) depending in shifted by imaginary half period argument into the formula (answer). To provide the reality of the solution we shall fix the parameters $a_i-\Delta$ in the permitted zones. The constants $C_i$ are computed by the formula (\ref{cccc}) at which $\nu$ means the coordinate of the curve (\ref{ctv4}). Consider further the Treibich Verdier potential \begin{equation} {\mathsf U}_5(x)=6\wp(x)+2\wp(x+\omega_2)+2\wp(x+\omega_3)+2e_1, \label{TV4} \end{equation} associated with four sheeted cover. The pontial is normalized according to (\ref{decomposition}). The associated spectral curve is of the form \begin{eqnarray}\nu^2&=&(\lambda+6e_2-3e_3)(\lambda+6e_3-3e_2) \nonumber\\ &\times&\left[\lambda^3+3e_1\lambda^2-(29e_2^2-22e_2e_3+29e_3^2)\lambda \right.\cr &+&\left.159(e_2^3+e_3^3)-51e_2e_3(e_2+e_3)\right] \label{ctv5} \end{eqnarray} The associated Hermite polynomials are given by the formula \begin{eqnarray*} {\mathcal F}(x,\lambda) &=&\lambda^2-(3\wp(x)+\wp(x+\omega_2))+\wp(x+\omega_3)+e_1) \lambda\\ &+&9\wp(x)(\wp(x)+\wp(x+\omega_2)+\wp(x+\omega_3 ) )+3\wp(x+\omega_2)\wp(x+\omega_3)\cr&+&3e_1(3\wp(x)+\wp(x+\omega_2)) +\wp(x+\omega_3))-\frac{39}{2}g_2+54e_1^2. \end{eqnarray*} The solution of the system results the substitution of these expressions to (\ref{answer}) as before, but this solution is blowing up. We remark, that since McKean, Moser and Airlault paper \cite{amm77} is well known, that all elliptic potentials of the Schr\"odinger equations and their isospectral transformation under the action of the KdV flow has the form, \begin{equation} {\mathsf U}(x)=2\sum_{i=1}^N\wp(x-x_i(t)),\label{elpot} \end{equation} The number $N$ is a positive integer $N>2$ (the number of ``particles") and the numbers $\boldsymbol{ x}=(x_1(t),\ldots,x_N(t))$ belongs to the locus ${\mathcal L}_N$, i.e., the geometrical position of the points given by the equations \begin{equation} {\mathcal L}_N=\left\{(\boldsymbol{ x});\sum_{i\neq j}\wp'(x_i(t)-x_j(t))=0,\; j=1,\ldots N\right\}.\label{locus} \end{equation} If the evolution of the particles $x_i$ over the locus is given by the equations, \[\frac{{\mathrm d} x_i}{ {\mathrm d} t}=6\sum_{j\neq i}\wp(x_i(t)-x_j(t)) \] then the potential (\ref{elpot}) is the elliptic solution to the KdV equation. Henseforth the elliptic potentials discuss can serve as imput for the isospectral deformation along the locus. Moreover these elliptic potential do not exhausted all the viriety of elliptic potential; we can mention here the elliptic potentials of Smirnov (\cite{sm89,sm94}) for which the shifts $x_i$ are not half periods. The involving of these objects to the subject can enlarge the classes of elliptic solutions to the system (\ref{manakov}) \section{Conclusions} In this paper we have described a family of elliptic solutions for the coupled nonlinear Schr\"dinger equations using Lax pair method and the general method of reduction of Abelian functions to elliptic functions. Our approach is systematic in the sense that special solutions (periodic, soliton etc.) are obtained in a unified way. We considered only the family of elliptic solutions associated with the integrable case $1:2:1$ of quartic potential, the approach developed can be applied to other integrable cases being enumerated in the introduction. In fiber optics applications of the periodic and quasi-periodic waves are of interest in optical transmission systems.
\section{Introduction} The orbit of Pluto has a number of unusual features. It has the highest eccentricity ($e_p=0.253$) and inclination ($i_p=17.1^\circ$) of any planet in the solar system. It crosses Neptune's orbit and hence is susceptible to strong perturbations during close encounters with that planet. However, close encounters do not occur because Pluto is locked into a 3:2 orbital resonance with Neptune, which ensures that conjunctions occur near Pluto's aphelion (\cite{CH65}). More precisely, the critical argument $3\lambda_p-2\lambda_n-\varpi_p$ librates around $180^\circ$ with a period of $1.99\times 10^4$ yr and an amplitude of $82^\circ$; here $\lambda_p$ and $\lambda_n$ are the mean longitudes of Pluto and Neptune and $\varpi_p$ is Pluto's longitude of perihelion. Other resonances are present with longer periods: for example, Pluto's argument of perihelion librates with an amplitude of $23^\circ$ and a period of 3.8 Myr. In part because of its rich set of resonances, Pluto's orbit is chaotic, although it exhibits no large-scale irregular behavior over Gyr timescales (see \cite{MW97} for a comprehensive review of Pluto's orbit). For reference, Pluto's semimajor axis and orbital period are $a_p=39.774{\rm\,AU}$ and $P_p=250.85$ yr. The most compelling explanation for Pluto's remarkable orbit was given by \nocite{M93,M95} Malhotra (1993, 1995). Malhotra argues that Pluto formed in a low-eccentricity, low-inclination orbit in the protoplanetary disk beyond Neptune. Subsequent gravitational scattering and ejection of planetesimals in the disk by all four giant planets caused Neptune's orbit to migrate outwards (\cite{FI84}). As its orbit expands, Neptune's orbital resonances sweep through the disk, first capturing Pluto into the 3:2 resonance and then pumping up its eccentricity. If Pluto's orbit was circular before capture, its present eccentricity implies that it was captured when Neptune's semimajor axis was 0.814 times its current value or $24.6{\rm\,AU}$ (eq. \ref{eq:renu}). This process may also excite Pluto's inclination although the details are less certain (\cite{M98}). Malhotra's argument predicts that most Kuiper belt objects with $30{\rm\,AU}\mathrel{\mathpalette\simov <} a\mathrel{\mathpalette\simov <} 50{\rm\,AU}$ should also be captured---and presently located---in Neptune resonances (\cite{M95}). This prediction has proved to be correct: of the $\sim 90$ Kuiper belt objects with reliable orbits as of 1999 January 1, over 30\% have semimajor axes within 1\% of the 3:2 resonance (although this number is exaggerated by observational selection effects). These objects have come to be called Plutinos, since they share the 3:2 resonance with Pluto (see \cite{MDL99} for a recent review of the Kuiper belt). Almost all studies of the dynamics of the Kuiper belt so far have neglected the gravitational influence of Pluto, because of its small mass ($M_p/M_\odot=7.40\times 10^{-9}$ for the Pluto-Charon system, \cite{S92,TW97}). However, like the Trojan asteroids and Jupiter, or the Saturn coorbital satellites Janus and Epimetheus, Pluto and the Plutinos share a common semimajor axis and hence even the weak gravitational force from Pluto can have a substantial influence on the longitude of a Plutino relative to Pluto. A crude illustration of the importance of Pluto's gravity is to note that the half-width of the 3:2 resonance, $(\Delta a/a)_{\rm res}\simeq 0.01$ for $0.2\mathrel{\mathpalette\simov <} e\mathrel{\mathpalette\simov <} 0.3$ (the maximum fractional amplitude of stable libration in semi-major axis, \cite{M96}), is only a few times larger than the Hill radius of Pluto, $(\Delta a/a)_{\rm H}=(M_p/3M_\odot)^{1/3}=0.0014$. The goal of this paper is to explore the dynamical interactions between Pluto and Plutinos and their consequences for the present structure of the Kuiper belt. Section 2 provides an approximate analytical description of the interactions, \S 3 describes the results of numerical orbit integrations, and \S 4 contains a discussion. \section{Analysis} We examine a simplified model solar system containing only the Sun and Neptune, with masses $M_\odot$ and $M_n$; we assume that Neptune's orbit is circular and neglect all orbital inclinations. We describe the motion of the Plutino using the canonical variables \begin{eqnarray} x_1=(GM_\odot a)^{1/2},\qquad\qquad & & y_1=\lambda, \nonumber \\ x_2=(GM_\odot a)^{1/2}[1-(1-e^2)^{1/2}],\qquad & & y_2=-\varpi, \label{eq:canvar} \end{eqnarray} where $a$ and $e$ are the semi-major axis and eccentricity, $\lambda$ and $\varpi$ are the mean longitude and longitude of perihelion. The same variables for Neptune or Pluto are denoted by adding a subscript ``n'' or ``p''. We consider only the resonant perturbations exerted by Neptune, which can only depend on angles as an even function of the combination $3y_1-2y_{n1}+y_2$. Thus the Hamiltonian of a Plutino may be written \begin{equation} H_0({\bf x},{\bf y},t)=H_K(x_1)+A({\bf x},3y_1-2y_{n1}+y_2)+B({\bf x},{\bf y},t), \end{equation} where $H_K(x_1)=-{\textstyle{1\over2}}(GM_\odot)^2/x_1^2$ is the Kepler Hamiltonian, $A$ is the resonant potential from Neptune, and $B({\bf x},{\bf y},t)$ is the potential from Pluto. The same Hamiltonian describes the motion of Pluto if we set $B=0$. Now impose a canonical transformation to new variables $({\bf J},{\bf w})$ defined by the generating function \begin{equation} S({\bf J},{\bf y},t)=J_1(3y_1-2y_{n1}+y_2)+\ffrac{1}{3}J_2(y_2-y_{p2}). \end{equation} Thus $w_1=\partial S/\partial J_1=3y_1-2y_{n1}+y_2$, $w_2=\partial S/\partial J_2=\frac{1}{3}(y_2-y_{p2})$, while $x_1=\partial S/\partial y_1=3J_1$, $x_2=\partial S/\partial y_2=J_1+\frac{1}{3}J_2$, and the new Hamiltonian is \begin{equation} H({\bf J},{\bf w},t)=H_0+{\partial S\over \partial t}= H_K(3J_1)-2\dot y_{n1}J_1-\ffrac{1}{3}\dot y_{p2}J_2+A(3J_1,J_1+\ffrac{1}{3}J_2,w_1)+B, \label{eq:ham} \end{equation} where $\dot y_{n1}$ is the mean motion of Neptune and $-\dot y_{p2}$ is the apsidal precession rate of Pluto. To proceed further we make use of the fact that $B/A=\hbox{O}(M_p/M_n)\ll 1$. Thus we can divide the motion of a Plutino into ``fast'' and ``slow'' parts. The fast motion is determined by the Kepler Hamiltonian $H_K$ and the resonant potential $A$ (this is the opposite of normal usage, where the resonant perturbations from Neptune are regarded as ``slow'' compared to non-resonant perturbations). The slow variations are caused by the Pluto potential $B$. First we examine the fast motion. We drop the potential $B$, so that $w_2$ is ignorable and $J_2=3x_2-x_1=(GM_\odot a)^{1/2}[2-3(1-e^2)^{1/2}]$ is a constant of the motion. Thus if Pluto was captured into resonance from a circular orbit with semimajor axis $a_i$, its present semimajor axis and eccentricity are related by \begin{equation} e^2=\ffrac{5}{9}-\ffrac{4}{9}(a_i/a)^{1/2}-\ffrac{1}{9}(a_i/a). \label{eq:renu} \end{equation} We write $J_1=J_{1r}+\Delta J_1$ where $J_{1r}$ is chosen to satisfy the resonance condition for the Kepler Hamiltonian, \begin{equation} 2\dot y_{n1}=3\dot y_1=3 {dH_K\over dx_1}\qquad\hbox{or} \qquad J_{1r}=\ffrac{1}{3}x_{1r}={(GM_\odot)^{2/3}\over (18\dot y_{n1})^{1/3}}. \label{eq:chosen} \end{equation} Since $A/H_K=\hbox{O}(M_n/M_\odot)\ll1$ we expect $|\Delta J_1|\ll J_{1r}$, so we can expand $H_K$ to second order in $\Delta J_1$; dropping unimportant constant terms the fast motion is determined by the Hamiltonian \begin{eqnarray} H_f(\Delta J_1,w_1) & = & \frac{9}{2}\left({d^2H_K\over dx_1^2}\right)_{x_{1r}} (\Delta J_1)^2+A(3J_{1r}+3\Delta J_1,J_{1r}+\Delta J_1+\ffrac{1}{3}J_2,w_1) \nonumber \\ & = & -{27\over 2a^2}(\Delta J_1)^2+A(3J_{1r}+3\Delta J_1,J_{1r}+\Delta J_1+\ffrac{1}{3}J_2,w_1). \label{eq:pend} \end{eqnarray} The Hamiltonian is autonomous and hence has a conserved energy $E_f=H_f$ and action $I=(2\pi)^{-1}\oint \Delta J_1 dw_1$. The motion is along the level surfaces of $H_f$ in the $(\Delta J_1,w_1)$ plane and typically consists of either libration ($w_1$ oscillates between fixed limits) or circulation ($w_1$ increases or decreases without reversing), just as in the case of the pendulum Hamiltonian. The stable equilibrium solutions (i.e. zero-amplitude libration) are given by \begin{equation} \Delta J_1={a^2\over 9}\left({\partial A\over\partial x_1}+\frac{1}{3}{\partial A\over\partial x_2}\right), \qquad {\partial A\over \partial w_1}=0, \qquad {\partial^2A\over \partial w_1^2}<0. \label{eq:eq} \end{equation} The slow motion is determined by averaging the Hamiltonian (\ref{eq:ham}) over the fast motion: \begin{equation} H_s(J_2,w_2,t)=E_f(J_2)-\ffrac{1}{3}\dot y_{p2}J_2+\langle B\rangle. \label{eq:hamslow} \end{equation} Here $\langle\cdot\rangle$ indicates an average over one period of the fast motion. The fast energy $E_f$ depends on $J_2$ through the constraint that the fast action $I$ is adiabatically invariant. \subsection{Solutions with zero-amplitude libration} The solutions to the fast and slow equations of motion are particularly simple in the case where the fast libration amplitude is zero for both Pluto and the Plutino. This approximation is not particularly realistic---the libration amplitude of Pluto is $82^\circ$---but illustrates the principal features of the Plutino motions. In this case the fast energy is \begin{equation} E_f=A(3J_{1r},J_{1r}+\ffrac{1}{3}J_2,w_{1r}) \end{equation} where $w_{1r}$ is the equilibrium angle given by equation (\ref{eq:eq}) and we have dropped much smaller terms that are O$(A^2)$. For simplicity we shall assume that there is only one stable equilibrium point, that is, one solution to equations (\ref{eq:eq}) for given $J_2$. At the equilibrium point the fast action is $I=0$, and the slow Hamiltonian (\ref{eq:hamslow}) is \begin{equation} H_s(J_2,w_2,t)=A(3J_{1r},J_{1r}+\ffrac{1}{3}J_2,w_{1r}) -\ffrac{1}{3}\dot y_{p2}J_2 + \langle B(3J_{1r},J_{1r}+\ffrac{1}{3}J_2,y_1,y_2,t)\rangle+\hbox{O}(A^2), \end{equation} where $y_1=\frac{1}{3}(w_{1r}+2y_{n1}-y_2)$. Since Pluto also is assumed to have zero libration amplitude, $y_{p1}=\frac{1}{3}(w_{1r}+2y_{n1}-y_{p2})$. Thus \begin{equation} 3(y_{p1}-y_1)=y_2-y_{p2}=3w_2 \qquad \hbox{mod}\,(2\pi); \label{eq:resrel} \end{equation} that is, the difference in longitude of perihelion between Pluto and the Plutino is three times the difference in mean longitude. The same result will hold true on average even if the libration amplitudes are non-zero. Now let us assume in addition that the eccentricities of both Pluto and the Plutino are small. Since their semi-major axes are the same, the gravitational potential from Pluto at the Plutino may be written \begin{equation} B=GM_p\left({{\bf r}\cdot{\bf r}_p\over|{\bf r}_p|^3}-{1\over|{\bf r}-{\bf r}_p|}\right) = {GM_p\over a}\left[\cos(y_1-y_{p1})-{1\over 2|\sin{\textstyle{1\over2}}(y_1-y_{p1})|} \right]. \end{equation} Using equation (\ref{eq:resrel}) this simplifies to \begin{equation} B={GM_p\over a}\left(\cos w_2-{1\over 2|\sin{\textstyle{1\over2}} w_2|} \right). \end{equation} Moreover \begin{equation} {dy_{p2}\over dt}={\partial A\over \partial x_2}(3J_{1r},J_{1r}+\ffrac{1}{3}J_{p2},w_{1r}) +\hbox{O}(A^2); \end{equation} thus the slow Hamiltonian can be rewritten as \begin{eqnarray} H_s(J_2,w_2)=A(3J_{1r},J_{1r}+\ffrac{1}{3}J_2,w_{1r}) -\ffrac{1}{3}J_2{\partial A\over\partial x_2}(3J_{1r},J_{1r}+\ffrac{1}{3}J_{p2},w_{1r}) \nonumber \\ +{GM_p\over a}\left(\cos w_2-{1\over 2|\sin{\textstyle{1\over2}} w_2|} \right) +\hbox{O}(A^2). \end{eqnarray} The interesting behaviour occurs when the actions $J_2$ and $J_{p2}$ are similar (i.e. the eccentricities of Pluto and the Plutino are similar), so we write $J_2=J_{p2}+\Delta J_2$ and expand $A$ to second order in $\Delta J_2$. Dropping unimportant constants and terms of O$(A^2)$ we have \begin{equation} H_s(\Delta J_2,w_2)=\frac{1}{18}\Delta J_2^2A_{22} (3J_{1r},J_{1r}+\ffrac{1}{3}J_{p2},w_{1r}) +{GM_p\over a}\left(\cos w_2-{1\over 2|\sin{\textstyle{1\over2}} w_2|} \right). \label{eq:tro} \end{equation} where $A_{22}=\partial^2A/\partial x_2^2$. This Hamiltonian is strongly reminiscent of the Hamiltonian for a test particle coorbiting with a satellite, \begin{equation} H_c(\Delta x,w)=-{3\over 2a^2}\Delta x^2 +{GM\over a}\left(\cos w-{1\over 2|\sin{\textstyle{1\over2}} w|} \right); \label{eq:hamc} \end{equation} here $\Delta x=x_1-x_{s1}$ and $w=\lambda_1-\lambda_{s1}$ are conjugate variables, and the subscript $s$ denotes orbital elements of the satellite. In this case the torques from the satellite lead to changes in semi-major axis; for a Plutino the semi-major axis is locked to Neptune's by the resonance, so torques from Pluto lead to changes in the eccentricity instead. When $A_{22}<0$ many of the features of orbits in the slow Hamiltonian (\ref{eq:tro}) follow immediately from the analogy with the Hamiltonian (\ref{eq:hamc}), which has been studied by many authors (e.g. Yoder et al. 1983, Namouni et al. 1999). The trajectories are determined by the level surfaces of the Hamiltonian. The equilibrium solutions correspond to the triangular Lagrange points in the coorbital case: $\Delta J_2=0$, $w_2=\pm 60^\circ$, $H_s=-{\textstyle{1\over2}} GM_p/a$; the eccentricity of the Plutino equals the eccentricity of Pluto, the mean longitude leads or lags by $60^\circ$, and the perihelia are $180^\circ$ apart. These solutions are maxima of the potential from Pluto. For smaller values of $H_s$, the orbits librate around the triangular points (``tadpole orbits''). Small-amplitude tadpole librations have frequency $\omega$ given by \begin{equation} \omega^2=-\frac{1}{4}A_{22}{GM_p\over a}. \label{eq:freq} \end{equation} The tadpole orbits merge at the separatrix orbit, $H_s=-\frac{3}{2}GM_p/a$; for this orbit the minimum separation is $w_{2,\rm min}=23.91^\circ$. Even smaller values of $H_s$ yield ``horseshoe'' orbits, with turning points at $\pm w_{2,\rm min}$ where $H_s=(GM_p/a)(\cos w_{2,\rm min}-{\textstyle{1\over2}}|\sin {\textstyle{1\over2}} w_{2,\rm min}|^{-1})$. For all tadpole and horseshoe orbits, the maximum and minimum values of $\Delta J_2$ occur at $w_2=\pm 60^\circ$, and are given by \begin{equation} \Delta J_2=\pm\left[{18\over A_{22}}\left(H_s+{GM_p\over 2a}\right)\right]^{1/2}. \label{eq:sep} \end{equation} Eventually the theory breaks down, when $w_{2,\rm min}$ is small enough that adiabatic invariance is no longer a valid approximation. \subsection{The resonant potential from Neptune} For quantitative applications we must evaluate the resonant Neptune potential $A({\bf x},w_1)$. For small eccentricities, the potential can be derived analytically, \begin{eqnarray} A({\bf x},w_1) &= & -{GM_n\over a}\bigg[{\textstyle{1\over2}} b_{1/2}^{(0)}(\alpha) +\ffrac{1}{8}e^2(2\alpha D+\alpha^2D^2)b_{1/2}^{(0)}(\alpha) + {\textstyle{1\over2}} e(5+\alpha D)b_{1/2}^{(2)}(\alpha)\cos w_1 \nonumber \\ & & +\ffrac{1}{8}e^2(104 +22\alpha D+\alpha^2D^2)b_{1/2}^{(4)}(\alpha) \cos 2w_1 +\hbox{O}(e^3)\bigg], \label{eq:anal} \end{eqnarray} where $\alpha=a_n/a<1$, $D=d/d\alpha$, \begin{equation} b_{1/2}^{(j)}(\alpha)={1\over\pi}\int_0^{2\pi} {d\phi \cos j\phi \over (1-2\alpha\cos\phi+\alpha^2)^{1/2}}\end{equation} is a Laplace coefficient, Neptune is assumed to be on a circular orbit, and inclinations are neglected. Unfortunately, in the present case the high eccentricity of the Plutino orbits makes this expansion invalid. However, we may determine $A({\bf x},w_1)$ numerically for given actions ${\bf x}$ by averaging the gravitational potential from Neptune over $y_2$ at fixed $w_1$, that is, \begin{equation} A({\bf x},w_1)=-{GM_n\over a}F({\bf x},w_1), \label{eq:resdef} \end{equation} where \begin{equation} F({\bf x},w_1)= {a\over 6\pi}\int_0^{6\pi} dy_2\left ({1\over |{\bf r}_n-{\bf r}|}-{{\bf r}\cdot{\bf r}_n\over |{\bf r}_n|^3}\right)_{{\bf x},w_1,y_2}+ \hbox{constant}; \label{eq:AF} \end{equation} the unimportant constant is chosen so that $A=0$ for circular orbits, i.e. $F(x_1,0,w_1)=0$. It can be shown analytically that the contribution from the second (indirect) term in the integrand vanishes. We shall also write \begin{equation} A_{22}={\partial^2 A\over \partial x_2^2}\equiv -{M_n\over M_\odot a^2} F_{22}({\bf x},w_1)\quad \hbox{where} \quad F_{22}({\bf x},w_1)=x_1^2{\partial^2F({\bf x},w_1)\over \partial x_2^2}. \label{eq:twotwo} \end{equation} Figure \ref{fig:contone}\ plots the contours of $F({\bf x},w_1)$ at the resonant semimajor axis $x_{1r}$, as obtained from equation (\ref{eq:AF}). The potential is singular for collision orbits, which for small eccentricity satisfy \begin{equation} \cos w_1={a-a_n\over ea}. \label{eq:coll} \end{equation} The conditions (\ref{eq:eq}) for stable zero-amplitude libration are satisfied if and only if $w_1=w_{1r}=\pi$ or $w_1=0$ and $e>1-a_n/a=0.237$. Figure \ref{fig:corot180}\ and \ref{fig:corot0}\ show examples of these two solutions, plotted in a reference frame corotating with Neptune. Orbits of the first kind are similar to Pluto's, although with smaller libration amplitude (compare Fig. 4 of \cite{MW97}). Orbits of the second kind (Fig. \ref{fig:corot0}) were discussed by \nocite{M96} Malhotra (1996), who calls them ``perihelion librators''. We shall not discuss these further, since they do not appear to form naturally during resonance capture of initially circular orbits; moreover for $e\mathrel{\mathpalette\simov >} 0.35$ they are likely to be unstable, since they cross Uranus's orbit and thus are subject to close encounters and collisions with that planet. Figure \ref{fig:approx}\ plots $F(x_{1r},x_2,\pi)$ and $F_{22}(x_{1r},x_2,\pi)$; at the eccentricity of Pluto, corresponding to $x_2/(GM_\odot a)^{1/2}=0.0325$, we have $F(x_{1r},x_2,\pi)=-0.313$ and $F_{22}(x_{1r},x_2,\pi)=79.3$. Thus, for example, the eccentricity oscillation in the separatrix orbit that marks the boundary between tadpole and horseshoe orbits has amplitude $\Delta e =0.007$ (eq. \ref{eq:sep}) and the period of libration of small tadpole orbits is $2\pi/\omega=9.1\times 10^{7}$ yr (eq. \ref{eq:freq}). For our purposes it is sufficient to work with the following numerical approximation to the resonant potential: \begin{equation} \widetilde F({\bf x},w_1)=-\frac{0.584+0.130e}{1+1.709e}\ln \left|1-4.222e\cos w_1\right|, \label{eq:F} \end{equation} where $x_2=x_{1r}[1-(1-e^2)^{1/2}]$. This approximation is chosen to match the resonant potential at the resonant semimajor axis $x_1=x_{1r}$; the dependence on the relative semimajor axes of Neptune and the Plutino is suppressed since the effects of this potential are only important near resonance. The logarithmic factor is chosen to reproduce the singularity in the resonant potential near the collision orbits defined approximately by equation (\ref{eq:coll}). The approximation formula also matches the analytic formula (\ref{eq:anal}) to O$(e^2)$ at $w_1=\pi$. Figure \ref{fig:conttwo}\ shows the contour plot analogous to Figure \ref{fig:conttwo}\ for the approximate resonant potential $\widetilde F$, and the triangles in Figure \ref{fig:approx}\ show $\widetilde F$ and $\widetilde F_{22}$. The agreement is very good, especially considering that errors are amplified by taking the two derivatives required to generate $\widetilde F_{22}$. \section{Numerical experiments} We follow the orbital evolution of Pluto and a Plutino in a simplified version of the Sun-Neptune-Pluto-Plutino four-body system that isolates the resonant potential from Neptune. Neptune is assumed to have a circular orbit that migrates outward according to the rule (\cite{M93}) \begin{equation} a_n(t)=a_f-\Delta a\exp(-t/\tau), \end{equation} where $a_f=30.17{\rm\,AU}$ is Neptune's present semimajor axis, $\Delta a=6{\rm\,AU}$, and $\tau=1.5$ Myr. Thus Neptune's initial semimajor axis is $24.17{\rm\,AU}$ and the initial location of the 3:2 orbital resonance is $31.67{\rm\,AU}$. The initial eccentricity of Pluto is taken to be zero and its initial semimajor axis is $33{\rm\,AU}$, as required so that its present eccentricity matches the observed value (eq. \ref{eq:renu}). We followed 160 test particles, with initial semimajor axes distributed uniformly in the range $[31{\rm\,AU},39{\rm\,AU}]$ and eccentricities distributed uniformly in the range $[0,0.03]$. The inclinations of Pluto and the test particles are chosen randomly in the range $[0,3^\circ]$ and their angular elements are chosen randomly from $[0,2\pi]$. Pluto and the test particles feel the resonant potential from Neptune, as defined by equations (\ref{eq:resdef}) and (\ref{eq:F}), but no other Neptune forces. The effects of the resonant Neptune potential on the orbital elements of Pluto and the test particles are followed using Lagrange's equations. The test particles do not influence Pluto or one another. However, they are subject to the gravitational potential from Pluto, \begin{equation} B({\bf x},{\bf y},t)=-GM_p\left({1\over |{\bf r}-{\bf r}_p|}-{{\bf r}\cdot{\bf r}_p\over|{\bf r}_p|^3}\right); \label{eq:B} \end{equation} the effects of this potential on the orbital elements of the test particles are followed using Gauss's equations. The evolution of Pluto and the test particles is followed for 0.45 Gyr or 10\% of the age of the solar system. \section{Results} Of the 160 test particles, all but 12 are captured into the 3:2 resonance with Neptune, in the sense that their final semimajor axes are close to $(3/2)^{2/3}a_n$ and their eccentricities are near the prediction of equation (\ref{eq:renu}), as shown in Figure \ref{fig:renu}. The 12 particles that are not captured lie inside the location of Neptune's 3:2 resonance at the start of the calculation, $(3/2)^{2/3}\times 24.17{\rm\,AU}=31.67{\rm\,AU}$, and would presumably be captured into other resonances if we used the full Neptune potential to work out their motion. We have verified this presumption by conducting shorter integrations ($1\times10^7 {\rm\,yr}$) using the same initial conditions but the complete Neptune potential. In this case all but 15 of the 160 particles were captured into the 3:2 resonance; the remainder were captured into the 4:3, 5:3 or 7:5 resonances. The behavior of the test particles in the 3:2 resonance (henceforth Plutinos) falls into the following broad classes: \begin{itemize} \item Tadpole orbits (5 particles): these have longitude difference $y_1-y_{p1}$ and differences in longitude of perihelion $y_{p2}-y_2$ that librate around the leading or trailing Lagrange point of Pluto (Figure \ref{fig:tadpole}). (Note that the libration center for the orbit in this Figure is $y_1-y_{p1}\simeq 100^\circ$, not $60^\circ$ as implied by the analysis in \S 2.1. This discrepancy arises because Pluto has a high eccentricity, while our analysis is only valid for near-circular orbits. Similarly, the perihelion difference librates around $\varpi-\varpi_p\simeq 300^\circ$, three times the difference in mean longitude as required by eq. \ref{eq:resrel}.) The tadpoles show no evidence of chaotic behavior or secular evolution over the length of our integration. The analysis in \S 2.2 suggests that the maximum eccentricity difference for these orbits is $\Delta e\simeq 0.007$; this requires in turn that their initial semimajor axes must have been close to Pluto's, as is seen to be the case in Figure \ref{fig:renu}. \item Horseshoe orbits (19 particles): the longitude difference oscillates around $180^\circ$, with jumps in the Plutino eccentricity at the extrema of the longitude oscillation, as predicted by the analysis of \S 2.1 (Figures \ref{fig:horseshoe1}, \ref{fig:horseshoe2}). The motion appears stable over the length of our integration although there are significant variations in semimajor axis oscillations during the course of the integration, and some horseshoes may evolve into transitional orbits over longer time intervals. \item Transitional orbits (43 particles): these show irregular behavior or transitions between libration and circulation of the longitude difference (Figure \ref{fig:trans}). When the longitude difference circulates, the particles are no longer protected from close encounters with Pluto. However, the particles remain in the 3:2 Neptune resonance in the sense that the resonant angle $w_1$ continues to librate. \item Doubly transitional orbits (2 particles): Like transitional orbits, these show libration-circulation transitions in the longitude difference, but in addition they show irregular behavior in the resonant angle $w_1$, leading eventually to a transition of $w_1$ from libration to circulation (Figure \ref{fig:dtrans}). Although only 2 particles in our simulation exhibit this behavior, a number of others show growing amplitude in the $w_1$ libration and will probably move into this class in less than the age of the solar system. Such orbits are normally short-lived since once $w_1$ circulates, they are no longer protected from close encounters with Neptune. \item Irregular circulating orbits (17 particles): the longitude difference circulates throughout the integration. Pluto induces irregular behavior (Figure \ref{fig:unaffect2}), but the Neptune resonance is preserved in the sense that $w_1$ continues to librate, at least over the span of our integration. \item Regular circulating orbits (62 particles): the longitude difference circulates throughout the integration, but the orbits appear fairly regular (Figure \ref{fig:unaffect1}). Generally, the orbits with larger eccentricity differences are more regular, because the encounter velocity with Pluto is higher so the perturbations from close encounters are smaller. \end{itemize} These classes represent a sequence in eccentricity difference: the typical eccentricity difference $|e-e_p|$ is smallest for tadpoles and largest for orbits unaffected by Pluto. Because the Plutino eccentricity is determined by the semimajor axis at the time of resonant capture (eq. \ref{eq:renu}), the classes also reflect the initial semimajor axes of the Plutinos: the tadpoles and horseshoes all have initial semimajor axes in the range 32.2{\rm\,AU}--34.2{\rm\,AU}\ (i.e. close to Pluto's initial semimajor axis of 33{\rm\,AU}). The transitional and irregular circulating Plutinos mostly have initial semimajor axes in the range 31.7{\rm\,AU}--36{\rm\,AU}, and the regular circulating Plutinos have initial semimajor axes concentrated in the range 35{\rm\,AU}--39{\rm\,AU}. \section{Discussion} Test particles captured into the 3:2 Neptune resonance (Plutinos) have a complex range of dynamical interactions with Pluto. The strength of the interaction depends on the difference in eccentricity between the test particle and Pluto, and thus on the difference in initial semi-major axis if the initial orbits were circular and capture occurred through outward migration of Neptune. Plutinos are stable only if the eccentricity difference $\Delta e$ is small ($\Delta e\mathrel{\mathpalette\simov <} 0.02$ from Figure \ref{fig:renu}), in which case the Plutinos librate on tadpole or horseshoe orbits; or if the eccentricity difference is large ($\Delta e\mathrel{\mathpalette\simov >} 0.06$), in which case the longitude difference circulates but relative velocity at encounter is high enough that Pluto has little effect. Unstable orbits at intermediate $\Delta e$ can be driven out of the 3:2 Neptune resonance by interactions with Pluto, and thereafter are short-lived because of close encounters with Neptune. Thus we expect that the population of Plutinos has decayed over time, although determining the survival fraction will require integrations over the lifetime of the solar system using the full Neptune potential. The long-term behavior of orbits in the 3:2 Neptune resonance is central to the origin of Jupiter-family comets. The usual explanation is that slow chaotic diffusion and collisional kicks drive Plutinos out of the 3:2 resonance, after which they are subjected to close encounters with the giant planets and eventually evolve into Jupiter-family comets (\cite{mo97}). Our results suggest that Pluto-induced evolution of Plutinos onto Neptune-crossing orbits may contribute to or even dominate the flux of Jupiter-family comets. Our results also enhance the motivation to obtain accurate orbits for Kuiper-belt objects, and fuel speculation that the formation of the Pluto-Charon binary may be linked to interactions between Pluto and Plutinos. This research was supported in part by NASA Grant NAG5-7310. We thank Matt Holman, Renu Malhotra, and Fathi Namouni for discussions and advice.
\subsection*{Introduction} As is known, there is presently no physical theory of the universe : the Standard Cosmological Model (SCM) seems to fail to explain many Observed Universe features because of the singularity problem. The old question why the universe is on average uniform and isotropic is still unanswered. The situation is aggravated, as new observational data on high-energy processes become available. In the suggested alternative cosmology the relativistic properties of a universe matter outside the Observed Universe are revealed. It gives us a clue for explaining major SCM difficulties. \section{CBR as a ``neo-ether''} In the SCM the whole Universe, which is the Observed Universe, represents a massive absolute reference frame matched with a Cosmic Background Radiation (SBR). The latter is even more perfect absolute reference system because it does not have peculiar velocity dispersion. Actually, this issue is similar to the one formulated in the question : why the Universe is so well ``tuned'' and ``aligned''. However, the ``neo-ether'' issue emphasizes more distinctly a confrontation of the SCM with conventional Physics in interpretation of the cosmological observations. If the Universe has such an inherent property as an absolute reference system then special and general relativistic theories must be redone. It is not easy to accept a CBR explanation of where the matter goes to in the expanding universe. But it is unacceptable to blame the general relativity theory for a decrease of a CBR energy density by one expansion factor faster than a total energy density of a massive corpuscular matter \cite{[1]}. The situation radically changes if one suggests that the Observed Universe is not the whole Universe but an ensemble of material objects in the Grand Universe. The latter is a multitude of different ensembles. Then a CBR as ether (an absolute reference frame) and other ``strange'' universe attributes become local features of the individual material system. This idea is put in a basis of the suggested alternative cosmological concept. \section{Antimatter issue} Antimatter is apparently present in the Observed Universe in small quantities while Physics shows no preference of matter over antimatter. If the whole universe is really baryon asymmetric then Physics must be deeply revised. The only way to save a baryon symmetry and a baryon charge conservation is again to suggest that the Observed (Home) Universe actually is for some reason a matter-made material system. One may call our Home Universe a representative of multitude of typical universes evenly made of either matter or antimatter and chaotically dispersed in a Grand Universe space. This is a continuation of the above idea of our Home Universe being a material system limited in volume. Thus, resolving the confrontation in baryon symmetry issue we can return to the characteristics of the whole Universe (Grand Universe, GU, for short) and to the concept of its evolution. \section{A Universe without absolute reference frame} An absence of the absolute reference system in the GU means a Lorentz invariance of a coordinate-momentum distribution function of a GU matter. It is known from statistics of a relativistic non-interacting gas \cite{[2]} that this function takes a form : $$F(x_1,x_2,x_3;p_1,p_2,p_3)dx_1dx_2dx_3dp_1dp_2dp_3 = Const\,dx_1dx_2dx_3dp_1dp_2dp_3 \eqno{(1)}$$ It characterizes a fully chaotic motion. In this idealized model an integral over a momentum distribution (1) diverges. It shows unlimited sky brightness (Olbers paradox). In practice a bolometer placed in GU space should measure a certain average energy density. It integrates radiation (matter flux) over a source distribution, which is a ``last scattering sphere''. Hence, a real momentum (energy) distribution must have a smooth cut--off of upper energies. This distribution is also a Lorentz--invariant because a last scattering sphere does not depend on a choice of a reference system. In terms of a velocity distribution of relativistic particles a special relativity theory gives a formula equivalent to (1), (see, for example, \cite{[2]}) : $$ F(\beta) d\beta = Const \gamma ^4 \beta ^2 d\beta \eqno{(2)}$$ for $0<\beta <\beta_{max}$ , where $\beta_{max}$ characterizes a cut--off parameter. The above relativistic distribution of GU matter is seen unchanged in any free reference system. All forms of matter are supposed to be subjected to this law. Hence, a thought observer can not distinguish between his states of relative motion in principle. Evidently, he can find any reference system among multitude of local ones, none of them being absolute. The important conclusion is that a GU matter can exist in a state of chaotic relativistic motion, which is seen identical for observers in any moving free reference system. In space with a matter in this state any kind of an ``absolute'' reference system allowing to detect any effect due to the relative motion does not exist. Now we can develop the alternative cosmological concept suggesting that the GU matter is evolving in a selfcreating manner being in a baryon symmetric steady state, which is found as a Lorentz--invariant (chaotic) relativistic motion. The GU space is thought to be of Minkovski's type. This is a scene for a cyclic evolution of a multitude of typical universes. Our Home Universe is one of them. \section{On a GU matter theory} The Grand Universe is an open system: there is no physical boundary. This is a primary reason for a GU matter being in a stationary state of a relativistic motion. Another important property of the GU system is a presence of two complementary entities, that is matter and antimatter. We may describe such a system in terms of a global conservation of sum of two kinds of energies : one is ``locked'' in a relativistic mass of matter and antimatter, the other is ``released'' in a matter-antimatter annihilation. Imagine for a moment that an initial state of the system is ``locked'' energy. Than an immediate process of energy release will start in a form of explosion (a GU version of ``inflation'' model). The other extreme case would be a state of full matter-antimatter annihilation. A ``pendulum'' starts going back through the process of pair creation restoring a selfsustained state of a relativistic motion. There is no energy dissipation, and the system acquires an equilibrium state in a form of continuing matter-antimatter annihilation and pair creation in parallel. Probably, this state may be described in terms of a maximal entropy of the GU system. It is clear now that we describe the idea of a generalized matter transport theory, which is different from Boltzmann's one first of all in above discussed two aspects : openness of the system, and matter annihilation/creation. Besides, we should introduce into this theory the basic laws of matter interactions, from gravity mechanics and physics, and nuclear physics, in particular. Boltzmann's statistics of gravitating objects is known as a mathematically very complex problem. Hence, we may think about model approximations on different levels. The most surprising what was found by author in a simple GU model is a statistical separation of matter and antimatter with a formation of an hierarchy structure, including evolving ``typical universes'' up to the size of our Home Universe. Evidently, only the smallest chances are given to a ``typical universe'' to reach a mature age. We expect that a development of a GU matter transport theory will give us a basis for a physical theory of Cosmology with a following development of specific models of a universe evolution, galaxy formation, cosmological nucleonic synthesis and others. It seems that the theory requires only one adjustable parameter : a matter density. We can not exclude even that a more general model may be developed with ``matter density long range waves''. In any case, the purpose of this theory would be solutions of the GU transport equation describing both a stationary Lorentz-invariant momentum distribution (discussed above) and an ``evolutionary function'' describing dynamics characteristics of matter structure, mass distribution of gravitationally linked systems (like evolving typical universes), in particular. For further discussing the main topic of this report we need a qualitative picture of the GU scenario of our Home Universe evolution. \section{Our Home Universe evolution} Any typical universe should start with some gravitating coagulant. Its growth to an embryonic or a mature state may be thought as the result of a probabilistic survival. Over a lifetime it continually interacts with the GU environment, some of the interactions being random ``catastrophic'' events. The mass distribution function mentioned must characterize this process that is actually fluctuations of the GU matter. This function evidently is a monotonous one rapidly decreasing. A ``tail'' of the function is due to the extremely rare and huge fluctuations what are typical universe formations. A random ``soft'' collision between them resulted in either further growth or a major annihilation. Typical universes as they reach a ``mature'' size become unaffected while capturing a ``small stuff'' (due to a good statistical averaging). For survival they compete with each other on a similar size level. In this scenario any typical universe is a gravitationally linked relativistic ensemble characterized by a momentum distribution similar to one of the GU matter but with a lower cut--off parameter. Eventually a growing universe becomes vulnerable to the collision with a quite smaller anti--counterpart. The criterion for this stage may be written roughly in a form: $$ G m / R c^2 (\gamma - 1) < 1 \eqno{(3)} $$ where : $m$ --- mass, $R$ --- mean radius, $\gamma$ --- effective Lorentz factor, $G$ --- gravitational constant, $c$ --- speed of light. Tracing back to redshifts $z\approx 10$ we can imagine our Home Universe being about one order smaller and about three orders denser than at present, what gives the criterion (3) a value close to the critical. If so, a mechanism triggering our early Home Universe to decay might be an abrupt mass drop due to an accidental collision with a smaller antimatter universe. As a result, we observe our present ``expanding'' Universe originated from some pre-expansion stage. Evidently, galaxies (many of them were formed at pre-expansion stage) are aligned in a Hubble's flow reflecting an initial relativistic distribution of a universe matter. So, chaos turned into an order. To complete the rough picture of Our Universe in the GU concept frame we should note that stars, galaxies and clusters are inner formations in a general GU structure hierarchy. {\it A luminous matter is expected to be only a small fraction of a total mass that provides for a self-sustained galaxy evolution.} Hence, a ``dark matter'', which is an ordinary matter, should be the dominant part in a typical universe. If so, it has to be in equilibrium with a thermal radiation. In our expanding Universe the observed 2.7 K CBR bears information on a mean surface temperature of dark matter bodies (dust clouds included). In accordance with the GU concept a mass distribution of dark matter is expected to be broad enough what makes a universe space quite transparent. Dark matter naturally participates in an ``expansion'' process. One can easily find that this concept perfectly explains all observed CBR features, a temperature decrease inversely proportional to the expansion factor, in particular. In a closed system like Our Universe, a thermal radiation performs an adiabatic expansion work because of a pressure gradient across the universe volume. Hence, there is no paradox with a ``missing energy''. Now we are ready to return to the main topic of this report : to give a qualitative explanation of high-energy cosmic corpuscular and gamma rays, gamma bursts, quasars and jetting objects in the frame of the suggested alternative Cosmology. \section{Cosmic rays} According to the GU concept a space outside typical universes is filled with a baryon symmetric matter in relativistic motion characterized by a Lorentz--invariant momentum distribution. The matter includes highly energetic gas and all sorts of macroscopic objects. Any typical universe evolves due to the active interaction with this outer GU background. Our Home Universe being also exposed to this background radiation must reveal products resulted from a transport of this radiation through the Universe medium. So called primary cosmic rays (its high--energy tail) should be actually a secondary radiation from interuniverse sources. In previous author's work \cite{[3]} this cosmic rays transport model is substantiated by numerical assessments. Cosmic rays researchers long time looked for natural mechanisms of acceleration. This is one of the old cosmological problems. The GU concept suggests an explanation of this phenomenon, paradoxically, as a contrary process of a moderation of primary ``inter--universe'' particles (both corpuscular and gamma quants) with much higher initial energies than observed. \section{Star explosions and gamma bursts} Gamma bursts discovered not long ago have been so far as enigmatic as cosmic rays. They happen to flash randomly and uniformly in the Universe space. Our simple explanation is as follows. The Universe space in fact contains an appreciable amount of antimatter in a form of material objects left from the moment of ``Big Bump''; some captured later from the outer space. As mentioned before, a ``dark matter'' partially consists of it. Some objects might be single stars. Hence, high--energy cosmic gamma rays should have a component due to the annihilation processes. A proper physical model is needed to distinguish this component from many other sources. But part of a radiation of an annihilation type is thought to be clearly observable due to its pulse character. There must be comparatively rare events of collision of matter and antimatter solid objects of big masses. In particular, the whole star may be involved in an annihilation event observed as a gamma burst of huge energy. The data already available on the identification of luminous ``disappearing'' objects as gamma burst sources. This explanation is given here as a hypothesis in the GU concept frame for a numerical test. \section{Quasars} We have to explain several quasar features : small size, big power density, age comparable with one of the Universe, power variability. According to the GU concept, quasars are ``afterglow'' resulting from the ``Big Bump'' what was a collision of our early Universe with some smaller typical universe made of antimatter (scenario discussed in previous sections). It must be the most catastrophic event in Our Universe evolution. A huge amount of matter has been annihilated with following release of energy in different forms of radiation, gamma radiation, in particular. A quasar ``engine'' should be a comparably small power--generating ensemble of material objects with a high surface--to--volume ratio. It keeps burning out in a non--exploding manner due to a continually incoming antimatter gas stream. One may explain its varying power by gas pressure instability. We think that so--called jetting objects have the same physical nature. A rough estimate of the quasar phenomenon in a simplified model \cite{[3]} shows a feasibility of matter--antimatter annihilation power generating mechanism. Detailed numerical analysis will be made after a development of the whole GU concept. \section{Conclusion} Now again the ``hot issues'' in a cosmological field are the old problems of the Standard Cosmological Model : is the Universe flat or open, and what is a true value of a Hubble's constant. As often happened in a science history, a resolution of a problem could be in finding that a problem is incorrectly formulated due to the wrong basic postulate. In the suggested alternative cosmology the answers to the above questions are : our Universe is neither open, nor flat, and an exact value of a Hubble's constant does not exist at all, for a peculiar velocity field has been initially created and further developed in both transversal and radial directions. The key element in a theory of the alternative cosmology must be the idea of a self--creation. A baryon symmetric world in Minkovski's space seems to be a perspective concept for a development of a cosmological theory having an explanatory and predictive power and being falsifiable. \subsection*{Acknowledgement} I want to express my gratitude to Professor Nikoly Rabotnov for interesting discussions of cosmological problems. I am also thankful to Professor Ryan Gieed for his interest and support of my work.
\section{Introduction} The ongoing microlensing observations towards the LMC and SMC have provided extremely puzzling results. On the one hand, analysis of the first two years of observations (Alcock et al.~1997a) suggest a halo composed of objects with mass $\sim 0.5{M_\odot}$ and a total mass in MACHOs out to 50 kpc of around $2.0 \times 10^{11}{M_\odot}$. On the other hand, producing such a halo requires extreme assumptions about star formation, galaxy formation, and the cosmic baryonic mass fraction. An attractive possibility is that the microlenses do not reside in the halo at all! Alternative suggested locations are the LMC halo (Kerins \& Evans 1999), the disk of the LMC itself (Sahu 1994), a warped and flaring Galactic disk (Evans et al.~1998), or an intervening population (Zhao 1998). Unfortunately, the low event rates, uncertainties in the Galactic model, and the velocity-mass-distance degeneracy in microlensing all conspire to make precise determinations of the MACHO parameters difficult. Over the next decade, second generation microlensing surveys, monitoring ten times the number of stars in the LMC will improve the overall statistics (and numbers of ``special'' events) considerably, allowing an unambiguous determination of the location of the microlenses. Even so, the paucity of usable lines of sight within our halo makes determination of the halo parameters such as the flattening or core radius very difficult. The Andromeda Galaxy (M31) provides a unique laboratory for probing the structure of galactic baryonic halos (Crotts 1992). Not only will the event rate be much higher than for LMC lensing, but it will be possible to probe a large variety of lines of sight across the disk and bulge and though the M31 halo. Furthermore, it provides another example of a bulge and halo which can be studied, entirely separate from the Galaxy. Recently, two collaborations, MEGA and AGAPE, have begun observations looking for microlensing in the stars of M31. Previous papers have made it clear that a substantial microlensing signal can be expected. In this paper we calculate, using realistic mass models, optical depth maps for M31. The results suggest that we should be able to definitively say whether M31 has a dark baryonic halo with only a few years or less of microlensing data. We also discuss how their variation with halo parameters may allow us to determine the M31 halo structure. This is particularly important in evaluating the level of resources that should be dedicated towards the ongoing observational efforts. Preliminary results suggest that the core radius and density profile power-law should be the easiest parameters to extract. The paper is organized in the following manner. In the next section we briefly discuss the M31 models we used. Following this we present optical depth maps for various halo models, discuss the microlensing backgrounds and finish with a quick discussion of the implications of the maps. \section{Modeling} Sources are taken to reside in a luminous two-component model of M31 consisting of an exponential disk and a bulge. The disk model is inclined at an angle of 77$^\circ$ and has a scale length of 5.8 kpc and a central surface brightness of $\mu_R = 20$ (Walterbos \& Kennicutt 1988). The bulge model is based on the ``small bulge'' of Kent (1989) with a central surface brightness of $\mu_R = 14$. This is an axisymmetric bulge with a roughly exp(-$r^{0.4}$) falloff in volume density with an effective radius of approximately 1 kpc and axis ratio, $c/a \sim 0.8$. Values of the bulge density are normalized to make $M_{bulge} = 4 \times 10^{10} {M_\odot}$. The predominant lens population is taken to be the M31 dark matter halo. We explore a parametrized set of M31 halo models. Each model halo is a cored ``isothermal sphere'' determined by three parameters: the flattening ($q$), the core radius ($r_c$) and the MACHO fraction ($f_b$): \begin{equation} \rho(x,y,z) = \frac{V_c(\infty)^2}{4 \pi G}\frac{e}{a^2q \sin^{-1} e} \frac{1}{x^2+y^2 + (z/q)^2 +a^2}, \end{equation} where a is the core radius, q is the x-z axis ratio, $e=\sqrt{1-q^2}$ and $V_c(\infty)=240$km/s is taken from observations of the M31 disk. In section 4 we briefly consider the optical depth due to other populations such as the bulge stars. More details of our modeling are given in Gyuk \& Crotts (1999) where in particular the velocity distributions (necessary for calculation of the microlensing {\em rate}) are discussed. These considerations do not affect the optical depths treated here. \section{Optical Depth Maps} \begin{figure} \epsfysize=10.0cm \centerline{ \rotate[r]{\epsfbox{tausliceminor.ps}}} \caption{Halo optical depth along the minor axis. The curves are: solid line -- q=0.3 core=1kpc, dotted line -- q=0.3 core= 5kpc, dashed line -- q=1.0 core=1.0kpc, and long dashed line -- q=1.0 core=5.0kpc } \label{tauminorslice} \end{figure} The classical microlensing optical depth is defined as the number of lenses within one Einstein radius of the source-observer line-of-sight (the microlensing tube): \begin{equation} \tau = \int_0^D \frac{\rho_{\rm halo}(d)}{M_{\rm lens}}\frac{4 G M_{\rm lens}}{c^2}\frac{(D-d)d}{D} \end{equation} Such a configuration is intended to correspond to a ``detectable magnification'' of at least a factor of $1.34$. Unfortunately, in the case of non-resolved stars (``pixel lensing'') we have typically \begin{equation} \pi \sigma^2 S_{M31} >> L_{*}. \end{equation} where $S_{M31}$ is the background surface brightness, $4\pi\sigma^2$ is the effective area of the seeing disk and $L_{*}$ is the luminosity of the source star. Thus it is by no means certain that a modest increase of $L_{*} \rightarrow 1.34 L_{*}$, as the lens passes within an Einstein radius, will be detectable. Furthermore, even for the events detected, measurement of the Einstein timescale $t_0$ is difficult. Thus measurement of the optical depth may be difficult. Nonetheless, advances have been made in constructing estimators of optical depth within highly crowded star fields (Gondolo 1999), which do not require the Einstein timescale for individual events, although they still require evaluation of the efficiency of the survey in question for events with various half maximum timescales. The errors on the derived optical depths will likely be larger than for the equivalent number of classical microlensing events. It is clear, however, that image image subtraction techniques (Tomaney \& Crotts 1996, Alcock et al.~1999a, b) can produce a higher event rate than conventional photometric monitoring. Thus one needs models of the optical depth, even if expressed only in terms of the cross-section for a factor 1.34 amplification in order to understand how microlensing across M31 will differ depending on the spatial distribution of microlensing masses in the halo and other populations. \begin{figure} \epsfysize=10.0cm \centerline{ \rotate[r]{\epsfbox{tauslicemajor.ps}}} \caption{Optical depth along a line parallel to the major axis and offset along the minor axis by -10$^\prime$ (towards the far side of the disk). The curves are: solid line -- q=0.3 core=1kpc, dotted line -- q=0.3 core= 5kpc, dashed line -- q=1.0 core=1.0kpc, and long dashed line -- q=1.0 core=5.0kpc } \label{taumajorslice} \end{figure} The above expression for the optical depth must be slightly amended to include the effects of the three-dimensional distribution of the source stars, especially of the bulge. We thus integrate the source density along the line of sight giving \begin{equation} \tau = \frac{\int_0^\infty \rho(S) \int_0^S \frac{\rho_{\rm halo}(s)}{M_{\rm lens}}\frac{4 G M_{\rm lens}}{c^2}\frac{(S-s)s}{S} ds dS} {\int_0^\infty \rho(S) dS} \end{equation} \begin{figure*} \centerline{ \hbox{ \hfill \vbox{ \epsfysize=10.0cm \rotate[r]{\epsfbox{tau_0.3_1.0.ps}} \epsfysize=10.0cm \rotate[r]{\epsfbox{tau_0.3_5.0.ps}}} \vbox{ \epsfysize=10.0cm \rotate[r]{\epsfbox{tau_1.0_1.0.ps}} \epsfysize=10.0cm \rotate[r]{\epsfbox{tau_1.0_5.0.ps}}} \hfill }} \caption{Contours of optical depth for halo models a) q=0.3 core=1.0, b) q=0.3 core=5.0, c) q=1.0 core=1.0 halo and d)q=1.0 core=5.0. Contours are from top to bottom: a) 2,3,4 and 5 $\times 10^{-6}$, b) 2,3 and 4 $\times 10^{-6}$, c) 1,2,3,4,5,6 and 7 $\times 10^{-6}$ and d) 1,2,3,4 and 5 $\times 10^{-6}$. For all models the dashed contour is 2.0$\times 10^{-6}$. } \label{taumodels} \end{figure*} The results of this calculation as a function of position for a variety of halo models are shown in Figure \ref{taumodels}. The most important attribute is the strong modulation of the optical depth from the near to far side of the M31 disk as was first remarked on by Crotts (1992). Near-side lines-of-sight have considerably less halo to penetrate and hence a lower optical depth. This can be seen nicely in Figure \ref{tauminorslice} where we plot the optical depth along the minor axis for the four models depicted in Figure \ref{taumodels}. While all models exhibit the strong variation from near to far, the fractional variation in $\tau$ across the minor axis is most pronounced for less flattened models, and changes in $\tau$ along the minor axis occur most rapidly for models with small core radii. This can be understood geometrically: in the limit of an extremely flattened halo the pathlength (and density run) through the halo is identical for locations equidistant from the center. Small core radii tend to make the central gradient steeper and produce a maximum at a distance along the minor axis comparable to the core size. This maximum is especially prominent in the flattened halos. Variations in core radii and flattening are also reflected in the run of optical depth along the major axis. In Figure \ref{taumajorslice} we show the optical depth along the major axis displaced by -10$^\prime$ on the minor axis. The gradients in the small core radii models are much larger than for large core radii. Asymptotically the flattened halos have a larger optical depth. \section{Background Lensing} Unfortunately, the M31 halo is not the only source of lenses. As mentioned above, the bulge stars can also serve as lenses. We show in Figure \ref{bulgelens} the optical depth contributed by the bulge lenses. The effect of the bulge lenses is highly concentrated towards the center. This is a mixed blessing. On the one hand the bulge contribution can thus be effectively removed by deleting the central few arcminutes of M31. Beyond a radius of 5 arcminutes, bulge lenses contribute negligibly to the overall optical depth. On the other hand the source densities are much higher in the central regions and thus we expect the bulk of our halo events to occur in these regions. We discuss this point in more detail in a forthcoming paper (Gyuk \& Crotts 1999). The bulge of M31 might easily serve as an interesting foil to the Galactic Bulge, which produces microlensing results which seem to require a special geometry relative to the observer, or other unexpected effects (Alcock et al.~1997b, Gould 1997). In addition to the M31 bulge lensing a uniform optical depth across the field will be contributed by the Galactic halo. This contribution will be of order $\sim 10^{-6}$ corresponding to a 40\% Galactic halo as suggested by the recent LMC microlensing results. Finally, disk self lensing will occur. The magnitude of the optical depth for this component will however be at least an order of magnitude lower than the expected halo or bulge contributions (Gould 1994) and hence is ignored in these calculations. \section{Discussion and Conclusions} The optical depth maps for M31 shown above exhibit a wealth of structure and clearly contain important information on the shape of the M31 halo. The most important of these information bearing characteristics is the asymmetry in the optical depth to the near and far sides of the M31 disk. A detection of strong variation in the optical depth from front to back will be a clear and unambiguous signal of M31's microlensing halo, perhaps due to baryons. No other lens population or contaminating background can produce this signal. However, the lack of a strong gradient should not be taken as conclusive proof that M31 does not have a halo. As discussed above, strong flattening or a large core radius can reduce or mask the gradient. Nevertheless, the halo should still be clearly indicated by the high microlensing rates observed outside the bulge region. In such a case, however, careful modeling of the experimental efficiency and control over the variable star contamination will be necessary to insure that the observed event are really microlensing. Further information about the structure of the M31 baryonic halo can be gleaned from the distribution of microlensing along the major axis. A strong maximum at the minor axis is expected for small core radii especially for spherical halos. The combination of the change in event rate both along the major and minor axis directions can in principle reveal both the core radius and flattening from a microlensing survey. How easily such parameters can be measured depends critically on the rate at which events can be detected, which we discuss in paper II of this series, along with estimates of the expected accuracy. Additionally, we will discuss strategies to optimize such surveys for measuring shape parameters. \begin{figure} \epsfysize=10.0cm \centerline{ \rotate[r]{\epsfbox{bulge_tau.ps}}} \caption{Contours of optical depth for the bulge self-lensing. Contours are, from the outside in: 1,2,3,4 and 5 $\times 10^{-6}$. Note that the region shown is half the dimensions of the maps of Figure 3.} \label{bulgelens} \end{figure}
\section{Introduction} Recent developments in the treatment of fermions in lattice gauge theory led to a better understanding of chiral symmetry not only in lattice theory [1]-[7]but possibly also in continuum theory[8]. These developments are based on a hermitian lattice Dirac operator $\gamma_{5}D$ which satisfies the so-called Ginsparg-Wilson relation[1] \begin{equation} \gamma_{5}D + D\gamma_{5} = aD\gamma_{5}D. \end{equation} An explicit example of the operator satisfying (1.1) and free of species doubling has been given by Neuberger[2]. The operator has also been discussed as a fixed point form of block transformations [3]. The relation (1.1) led to the interesting analyses of the notion of index in lattice gauge theory[4]-[9]. Here $\gamma_{5}$ is a hermitian chiral Dirac matrix. The index relation is generally written as [4][5] \begin{equation} Tr \gamma_{5}(1 -\frac{1}{2}aD) = n_{+} - n_{-} \end{equation} which is confirmed by [8] \begin{eqnarray} Tr [ \gamma_{5}(1-\frac{1}{2}aD)]&=& \sum_{n}\{\phi^{\dagger}_{n}\gamma_{5}\phi_{n} - \frac{1}{2}\phi^{\dagger}_{n}\gamma_{5}aD\phi_{n}\}\nonumber\\ &=& \sum_{ \lambda_{n}=0}\phi^{\dagger}_{n}\gamma_{5}\phi_{n} + \sum_{ \lambda_{n}\neq 0}\phi^{\dagger}_{n}\gamma_{5}\phi_{n} - \sum_{n}\frac{1}{2}a\lambda_{n}\phi^{\dagger}_{n}\phi_{n}\nonumber\\ &=&\sum_{ \lambda_{n}=0}\phi^{\dagger}_{n}\gamma_{5}\phi_{n}\nonumber\\ &=& n_{+} - n_{-} = index \end{eqnarray} where $n_{\pm}$ stand for the number of normalizable zero modes in \begin{equation} \gamma_{5}D\phi_{n}=\lambda_{n}\phi_{n} \end{equation} for the {\em hermitian} operator $\gamma_{5}D$ with simultaneous eigenvalues $\gamma_{5}\phi_{n}= \pm \phi_{n}$. We also used the relation \begin{equation} \phi^{\dagger}_{n}\gamma_{5}\phi_{n} = \frac{a}{2}\lambda_{n}\phi^{\dagger}_{n}\phi_{n} = \frac{a}{2}\lambda_{n} \end{equation} for $\lambda_{n}\neq 0$, which is derived by sandwiching the relation(1.1) by $\phi^{\dagger}_{n}\gamma_{5}$ and $\phi_{n}$. It should be emphasized that the relation (1.3) is derived {\em without} using $Tr \gamma_{5}=0$. The inner product $\phi^{\dagger}_{n}\phi_{n} = (\phi_{n},\phi_{n}) \equiv \sum_{x} a^{4}\phi^{\star}_{n}(x)\phi_{n}(x)$ is defined by summing over all the lattice points, which are not explicitly written in $\phi_{n}$. See Appendix for further notational details. An advantage of gauge theory defined on a finite lattice is that one can analyze some subtle aspects of chiral symmetry in continuum theory in a well-defined finite setting. The purpose of the present note is to study some of those aspects of chiral symmetry in the hope that this analysis also deepens our understanding of lattice regularization. In the path integral treatment of chiral anomaly in continuum, the relation \begin{equation} Tr \gamma_{5} = n_{+} - n_{-} \end{equation} in a suitably regularized sense plays a fundamental role[10][11]. On the other hand, it is expected that the relation \begin{equation} Tr \gamma_{5} = 0 \end{equation} holds on a finite lattice. As Chiu pointed out[12], this relation (1.7) leads to an interesting {\em constraint} \begin{equation} Tr \gamma_{5} = n_{+}- n_{-} + N_{+} - N_{-} = 0 \end{equation} where $N_{\pm}$ stand for the number of eigenstates $\gamma_{5}D\phi_{n}=\pm (2/a)\phi_{n}$ with $\gamma_{5}\phi_{n}= \pm \phi_{n}$, respectively. It is important to recognize that $Tr \gamma_{5} = 0$ means that this relation holds for {\em any} sensible basis set with any background gauge field in a given theory, which may be used to define the trace. Consequently, the seemingly trivial relation $Tr \gamma_{5} = 0$ in fact carries important physical information. In this note we show that $ Tr \gamma_{5} = 0$ implies the inevitable contribution from unphysical (would-be) species doublers in lattice theory or an unphysical bosonic spinor in Pauli-Villars regularization. In other words, $Tr \gamma_{5} = 0$ cannot hold in the physical Hilbert space consisting of physical states only, and the continuum limit of $ Tr \gamma_{5} = 0$ is not defined consistently, as is seen in (1.8). It is shown that the failure of the decoupling of heavy fermions in the anomaly calculation is crucial to understand the consistency of the customary lattice calculation of anomaly where $ Tr \gamma_{5} = 0$ is used. ( The continuum limit in this paper stands for the so-called ``naive'' continuum limit with $a\rightarrow 0$, and the lattice size is gradually extended to infinity for any finite $a$ in the process of taking the limit $a\rightarrow 0$.) We then discuss the possible implications of our analysis on the treatment of chiral anomalies in continuum theory. We also briefly comment on an analogous phenomenon in the analysis of photon phase operator, where the notion of index plays a crucial role. \section{Consistency of the relation $Tr \gamma_{5} = 0$ } In the previous section we have seen that the consistency of the relation $Tr \gamma_{5} = 0$ requires the presence of the $N_{\pm}$ states for an operator $\gamma_{5}D$ satisfying (1.1) on a finite lattice. We thus want to analyze the nature of the $N_{\pm}$ states in more detail. For this purpose, we start with the conventional Wilson operator $D_{W}$ \begin{eqnarray} D_{W}(n,m)&\equiv&i\gamma^{\mu}C_{\mu}(n,m) + B(n,m) -\frac{1}{a}m_{0}\delta_{n,m},\nonumber\\ C_{\mu}(n,m)&=&\frac{1}{2a}[\delta_{m+\mu,n}U_{\mu}(m) - \delta_{m,n+\mu}U^{\dagger}_{\mu}(n)],\nonumber\\ B(n,m)&=&\frac{r}{2a}\sum_{\mu}[2\delta_{n,m}-\delta_{m+\mu,n}U_{\mu}(m) -\delta_{m,n+\mu}U^{\dagger}_{\mu}(n)],\nonumber\\ U_{\mu}(m)&=& \exp [iagA_{\mu}(m)], \end{eqnarray} where we added a constant mass term to $D_{W}$ for later convenience. Our matrix convention is that $\gamma^{\mu}$ are anti-hermitian, $(\gamma^{\mu})^{\dagger} = - \gamma^{\mu}$, and thus $\not \!\! C\equiv \gamma^{\mu}C_{\mu}(n,m)$ is hermitian \begin{equation} \not \!\! C ^{\dagger} = \not \!\! C. \end{equation} Since the operator $\not \!\! C$ forms the basis for any fermion operator on the lattice, we start with the analysis of $\not \!\! C$. \subsection{Operator $\not \!\! C$ and $Tr \gamma_{5} = 0$} It was noted elsewhere[8] that $Tr \gamma_{5} = 0$ implies the species doubling for the operator $\not \!\! C$. The basic reasoning is based on the index relation \begin{equation} dim\ ker\ (\frac{1- \gamma_{5}}{2})\not \!\! C(\frac{1+\gamma_{5}}{2}) - dim\ ker\ (\frac{1+\gamma_{5}}{2})\not \!\! C(\frac{1- \gamma_{5}}{2})=0 \end{equation} where we understand $(\frac{1- \gamma_{5}}{2})\not \!\! C(\frac{1+\gamma_{5}}{2})$ as standing for the two-component operator $b$ in \begin{eqnarray} \not \!\! C&=&\left(\begin{array}{cc} 0&b^{\dagger}\\ b&0 \end{array}\right). \end{eqnarray} This form of $\not \!\! C$ is deduced by noting $\not \!\! C^{\dagger} = \not \!\! C$ and $\gamma_{5}\not \!\! C + \not \!\! C\gamma_{5}=0$ in the representation where $\gamma_{5} $ is diagonal. The operator $b(m,n)$ projects a two-component spinor on a finite lattice to another two-component spinor on the same lattice , and thus it is a square matrix in the coordinate representation. For a general finite dimensional square matrix $M$, the index theorem $dim\ ker\ M - dim\ ker\ M^{\dagger} = 0$ holds[8], where $dim\ ker\ M $, for example, stands for the number of normalizable modes in $Mu_{n}=0$. In the present context, $dim\ ker\ (\frac{1- \gamma_{5}}{2})\not \!\! C(\frac{1+\gamma_{5}}{2}) = dim\ ker\ b $ stands for the number of normalizable zero modes in \begin{equation} \not \!\! C\phi_{n}= 0 \end{equation} with $(\frac{1+\gamma_{5}}{2})\phi_{n}= \phi_{n}$. Thus the index relation (2.3) shows that possible zero modes with $\gamma_{5}\phi_{n}= \pm \phi_{n}$ are always paired. The eigenstates with non-zero eigenvalues in \begin{equation} \not \!\! C\phi_{n}= \lambda_{n}\phi_{n} \end{equation} give a vanishing contribution to the trace $ Tr \gamma_{5}$ since \begin{equation} \phi_{n}^{\dagger}\gamma_{5}\phi_{n} = 0 \end{equation} by noting $\gamma_{5}\not \!\! C+ \not \!\! C\gamma_{5}=0$. The index relation (2.3) is thus equivalent to $ Tr \gamma_{5}=0$. If one recalls the Atiyah-Singer index theorem[10][13] written in the same notation as (2.3) \begin{equation} dim\ ker\ (\frac{1- \gamma_{5}}{2})\not \!\! D(\frac{1+\gamma_{5}}{2}) - dim\ ker\ (\frac{1+\gamma_{5}}{2})\not \!\! D(\frac{1- \gamma_{5}}{2})=\nu \end{equation} where $\nu$ stands for the Pontryagin index (i.e., an integral of anomaly) and $\not \!\! D \equiv \gamma^{\mu} (\partial_{\mu} - igA_{\mu})$, one sees that a smooth continuum limit of the lattice index relation (2.3) for a general background gauge field configuration is {\em inconsistent} with the absence of species doublers. In the present $\not \!\! C$, a very explicit construction of species doublers is known. For a square lattice one can explicitly show that the simplest lattice fermion action \begin{equation} S = \bar{\psi}i\not \!\! C\psi \end{equation} is invariant under the transformation[14] \begin{equation} \psi^{\prime}= {\cal T}\psi,\ \bar{\psi}^{\prime}= \bar{\psi}{\cal T}^{-1} \end{equation} where ${\cal T}$ stands for any one of the following 16 operators \begin{equation} 1,\ T_{1}T_{2},\ T_{1}T_{3},\ T_{1}T_{4},\ T_{2}T_{3},\ T_{2}T_{4},\ T_{3}T_{4},\ T_{1}T_{2}T_{3}T_{4}, \end{equation} and \begin{equation} T_{1},\ T_{2},\ T_{3},\ T_{4},\ T_{1}T_{2}T_{3},\ T_{2}T_{3}T_{4},\ T_{3}T_{4}T_{1},\ T_{4}T_{1}T_{2}. \end{equation} The operators $T_{\mu}$ are defined by \begin{equation} T_{\mu}\equiv \gamma_{\mu}\gamma_{5}\exp {(i\pi x^{\mu}/a)} \end{equation} and satisfy the relation \begin{equation} T_{\mu}T_{\nu} + T_{\nu}T_{\mu}=2\delta_{\mu\nu} \end{equation} with $T_{\mu}^{\dagger} = T_{\mu} = T^{-1}_{\mu}$ for anti-hermitian $\gamma_{\mu}$. We denote the 16 operators by ${\cal T}_{n}, \ \ n=0\sim 15$, in the following with ${\cal T}_{0}=1$. By recalling that the operator $T_{\mu}$ adds the momentum $\pi/a$ to the fermion momentum $k_{\mu}$, we cover the entire Brillouin zone \begin{equation} - \frac{\pi}{2a} \leq k_{\mu} < \frac{3\pi}{2a} \end{equation} by the operation (2.10) starting with the free fermion defined in \begin{equation} - \frac{\pi}{2a} \leq k_{\mu} < \frac{\pi}{2a}. \end{equation} The operators in (2.11) commute with $\gamma_{5}$, whereas those in (2.12) anti-commute with $\gamma_{5}$ and thus change the sign of chiral charge, reproducing the 15 species doublers with correct chiral charge assignment; $\sum_{n=0}^{15}(-1)^{n}\gamma_{5} =0$. In a smooth continuum limit, the operaor $\not \!\! C$ produces $\not \!\! D$ for each species doubler with alternating chiral charge. The relation $Tr \gamma_{5} = 0$ or (2.3) for the operator $\not \!\! C$ is consistent for any background gauge field because of the presence of these species doublers, which are degenerate with the physical species in the present case. \subsection{Wilson operator $D_{W}$ and $Tr\gamma_{5} = 0$} The consistency of $Tr\gamma_{5} = 0$ is analyzed by means of topological properties which are specified by (2.8) and thus it is best described in the nearly continuum limit. To be more precise, one may define the near continuum configurations by the momentum $k_{\mu}$ carried by the fermion \begin{equation} - \frac{\pi}{2a}\epsilon \leq k_{\mu} \leq \frac{\pi}{2a}\epsilon \end{equation} for sufficiently small $a$ and $\epsilon$ combined with the operation ${\cal T}_{n}$ in (2.11) and (2.12). To identify each species doubler clearly in the near continuum configurations, we also keep $r/a$ and $m_{0}/a$ finite for $a\rightarrow$ small [14], and the gauge fields are assumed to be sufficiently smooth. For these configurations, we can approximate the operator $D_{W}$ by \begin{equation} D_{W}= i\not \!\! D + M_{n} + O(\epsilon^{2}) + O(agA_{\mu}) \end{equation} for each species doubler, where the mass parameters $M_{n}$ stand for $M_{0}= - \frac{m_{0}}{a}$ and one of \begin{eqnarray} &&\frac{2r}{a}-\frac{m_{0}}{a},\ \ (4,-1);\ \ \ \frac{4r}{a}-\frac{m_{0}}{a},\ \ (6,1)\nonumber\\ &&\frac{6r}{a}-\frac{m_{0}}{a},\ \ (4,-1);\ \ \ \frac{8r}{a}-\frac{m_{0}}{a},\ \ (1,1) \end{eqnarray} for $n=1\sim 15$. Here we denoted ( multiplicity, chiral charge ) in the bracket for species doublers. In (2.18) we used the relation valid for the configurations (2.17), for example, \begin{eqnarray} D_{W}(k) &=& \sum_{\mu}\gamma^{\mu}\frac{\sin ak_{\mu}}{a} + \frac{r}{a}\sum_{\mu}(1 - \cos ak_{\mu}) - \frac{m_{0}}{a}\nonumber\\ &=& \gamma^{\mu}k_{\mu}( 1 + O(\epsilon^{2})) + \frac{r}{a} O(\epsilon^{2}) - \frac{m_{0}}{a} \end{eqnarray} in the momentum representation with vanishing gauge field. In these near continuum configurations, the topological properties are specified by the operator $\not \!\! D$ in $D_{W}$. We can thus evaluate $Tr\gamma_{5}$ by using the basis set defined by \begin{equation} \not \!\! D\phi_{n} = \lambda_{n}\phi_{n} \end{equation} which formally {\em diagonalize} the effective operator $D_{W}$ in (2.18) describing the low-energy excitations of each species doubler. We then obtain \begin{equation} Tr \gamma_{5} = \sum_{n=0}^{15}(-1)^{n}\lim_{L\rightarrow large}\sum_{l=1}^{L}\phi_{l}^{\dagger}\gamma_{5}\phi_{l}=0 \end{equation} where $\phi_{l}^{\dagger}\gamma_{5}\phi_{l}=0$ for $\lambda_{l}\neq 0$ because of $\gamma_{5}\not \!\! D + \not \!\! D\gamma_{5}=0$, and (2.22) states the cancellation of zero- mode contributions $\sum_{\lambda_{l}=0}\phi_{l}^{\dagger}\gamma_{5}\phi_{l}$ among various species. We are assuming that our near continuum configurations (2.18) are accurate in the treatment of these zero modes. An argument to support our identification of the near continuum configurations will be given in the next sub-section. $Tr\gamma_{5}=0$ is thus consistent even for a topologically non-trivial gauge background because of the presence of the would-be species doublers. This property is related to the well-known fact that one can safely ignore the Jacobian factor for global chiral transformation $\delta\psi = i\epsilon\gamma_{5}\psi$ and $\delta\bar{\psi}= \bar{\psi}i\epsilon\gamma_{5}$ for the theory defined by $S=\bar{\psi}D_{W}\psi$. \subsection{Overlap Dirac operator and $Tr\gamma_{5} = 0$} The operator $D$ introduced by Neuberger[2] , which satisfies the relation (1.1), has an explicit expression \begin{equation} aD= 1 - \gamma_{5}\frac{H}{\sqrt{H^{2}}} =1 + D_{W}\frac{1}{\sqrt{D_{W}^{\dagger}D_{W}}} \end{equation} where $D_{W}=-\gamma_{5} H$ is the Wilson operator. For the near continuum configurations specified above in (2.17), one can approximate \begin{eqnarray} D&=& \sum_{n=0}^{15}(1/a){[}1 + (i\not \!\! D + M_{n})\frac{1}{\sqrt{\not \!\! D^{2} + M_{n}^{2}}}]|n\rangle\langle n|,\nonumber\\ \gamma_{5}D&=& \sum_{n=0}^{15}(-1)^{n}\gamma_{5}(1/a){[}1 + (i\not \!\! D + M_{n})\frac{1}{\sqrt{\not \!\! D^{2} + M_{n}^{2}}}]|n\rangle\langle n|,\nonumber\\ \gamma_{5}&=& \sum_{n=0}^{15}(-1)^{n}\gamma_{5}|n\rangle\langle n|. \end{eqnarray} Here we explicitly write the projection $|n\rangle\langle n|$ for each species doubler. The operators in (2.24) preserve the Ginsparg-Wilson relation (1.1). We can again use the basis set in (2.21), which formally diagonalize the basic operator $D$ in (2.24), to define the trace operation. We thus obtain \begin{equation} Tr \gamma_{5} = \sum_{n=0}^{15}(-1)^{n}\lim_{L\rightarrow large}\sum_{l=1}^{L}\phi_{l}^{\dagger}\gamma_{5}\phi_{l}=0 \end{equation} by assuming that our effective operators (2.24) are accurate in describing the excitations near the zero modes $\not \!\! D\phi_{l}=0$ , which are relevant for topological considerations. Again the presence of the would-be species doublers makes the relation $Tr \gamma_{5}=0 $ consistent for any topologically non-trivial background gauge field. A justification of our effective description (2.24) will be given later. The above expression of $D$ also shows that \begin{eqnarray} D\phi_{l}& =& 0, \nonumber\\ D\phi_{l}&=& \frac{2}{a}\phi_{l} \end{eqnarray} for the physical species and the unphysical species doublers, respectively, if one uses the zero-modes $\not \!\! D\phi_{l}=0$. Note that $M_{0}<0$ and the rest of $M_{n}>0$ in (2.19) and (2.24) [2]. We also note that $\phi_{l}$ can be a simultaneous eigenstate of $\gamma_{5}$ only for $\not \!\! D\phi_{l}=0$. Namely, the $N_{\pm}$ states with the eigenvalue $2/a$ in fact correspond to topological excitations associated with species doublers; this means that the multiplicities of these $N_{\pm}$ are quite high due to the 15 species doublers, although they satisfy the sum rule $n_{+} + N_{+} = n_{-} + N_{-}$. This sum rule is a direct consequence of (2.25) and (2.26) by noting that $\phi_{l}^{\dagger}\gamma_{5}\phi_{l}= 0$ for $\lambda_{l}\neq 0$. The calculation of the index (1.2) may proceed as \begin{eqnarray} Tr \gamma_{5}(1 - \frac{a}{2}D) &=& - Tr\gamma_{5}\frac{a}{2}D\nonumber\\ &=& - \frac{1}{2} Tr \sum_{n=0}^{15}(-1)^{n}\gamma_{5}{[}1 + (i\not \!\! D + M_{n})\frac{1}{\sqrt{\not \!\! D^{2} + M_{n}^{2}}}]\nonumber\\ &=& - \frac{1}{2} Tr \sum_{n=0}^{15}(-1)^{n}\gamma_{5}(i\not \!\! D + M_{n})\frac{1}{\sqrt{\not \!\! D^{2} + M_{n}^{2}}}\nonumber\\ &=& - \frac{1}{2} \sum_{n=0}^{15}(-1)^{n}\sum_{l}\phi_{l}^{\dagger}\gamma_{5} M_{n}\frac{1}{\sqrt{\not \!\! D^{2} + M_{n}^{2}}}\phi_{l}\nonumber\\ &=& - \frac{1}{2} \sum_{n=0}^{15}(-1)^{n}M_{n}\frac{1}{\sqrt{M_{n}^{2}}} \sum_{\lambda_{l}=0}\phi_{l}^{\dagger}\gamma_{5}\phi_{l}\nonumber\\ &=& - \frac{1}{2} \sum_{n=0}^{15}(-1)^{n}M_{n}\frac{1}{\sqrt{M_{n}^{2}}}(n_{+} - n_{-})\nonumber\\ &=& - \frac{1}{2}(-1 + \sum_{n=1}^{15}(-1)^{n})(n_{+} - n_{-}) = n_{+} - n_{-} \end{eqnarray} where we used $\gamma_{5}\not \!\! D + \not \!\! D\gamma_{5}=0$ and the fact that $\phi_{l}^{\dagger}\gamma_{5}\phi_{l}= 0$ for $\lambda_{l}\neq 0$ in $\not \!\! D\phi_{l}= \lambda_{l}\phi_{l}$. We also used the fact that $M_{0}<0$ and $M_{n}>0$ for $n=1\sim 15$ [2]. The index (2.27) is defined for $ \not \!\! D$ while the index (1.3) is defined for $\gamma_{5}D$, and both agree with the Pontryagin index as is seen in (2.30) below. In the above calculation (2.27), we used the relation $Tr \gamma_{5}= 0$ twice: In the second line, this relation requires the presence of the physical species as well as the species doublers. As a result, we have the contribution to the final index from both of the physical species and 15 species doublers, although the species doublers with $\lambda_{l}= 2/a$ should saturate the index (and anomaly) in the expression[12] \begin{equation} - Tr (a/2)\gamma_{5}D = n_{+} - n_{-} \end{equation} as is noted in (A.10) in Appendix. Our analysis of the global topological property on the basis of effective operators (2.24) is thus consistent. The above calculational scheme of index (2.27) in fact corresponds to the evaluation of the local index (i.e., anomaly) performed in Ref.[8]. By using the plane wave basis , one has (in the limit $a\rightarrow 0$ with $r/a$ and $m_{0}/a$ kept fixed ) \begin{eqnarray} &&tr \gamma_{5}(1 - \frac{a}{2}D)(x)\nonumber\\ &=& - \frac{1}{2} \sum_{n=0}^{15}(-1)^{n}tr \int^{\infty}_{-\infty} \frac{d^{4}k}{(2\pi)^{4}} e^{-ikx}\gamma_{5}(i\not \!\! D + M_{n})\frac{1}{\sqrt{\not \!\! D^{2} + M_{n}^{2}}}e^{ikx}\nonumber\\ &=& \frac{1}{2}tr \int^{\infty}_{-\infty} \frac{d^{4}k}{(2\pi)^{4}} e^{-ikx}\gamma_{5}\frac{1}{\sqrt{\not \!\! D^{2}/M_{0}^{2} + 1}}e^{ikx}\nonumber\\ &&- \frac{1}{2}\sum_{n=1}^{15}(-1)^{n}tr \int^{\infty}_{-\infty} \frac{d^{4}k}{(2\pi)^{4}} e^{-ikx}\gamma_{5}\frac{1}{\sqrt{\not \!\! D^{2}/M_{n}^{2}+1}}e^{ikx} \end{eqnarray} which gives rise to the anomaly for all $|M_{n}| \rightarrow \infty$ in the continuum limit: \begin{eqnarray} tr \gamma_{5}(1 - \frac{a}{2}D)(x) &=& \lim_{M\rightarrow \infty}tr \int^{\infty}_{-\infty} \frac{d^{4}k}{(2\pi)^{4}} e^{-ikx}\gamma_{5}f(\frac{\not \!\! D^{2}}{M^{2}})e^{ikx}\nonumber\\ &=& \frac{g^{2}}{32\pi^{2}}tr \epsilon^{\mu\nu\alpha\beta}F_{\mu\nu}F_{\alpha\beta}. \end{eqnarray} Here we defined $f(x)= 1/\sqrt{x+1}$ which satisfies \begin{eqnarray} && f(0)=1,\ \ f(\infty)=0,\nonumber\\ && f^{\prime}(x)x|_{x=0} = f^{\prime}(x)x|_{x=\infty} = 0. \end{eqnarray} The right-hand side of (2.30) is known to be independent of the choice of $f(x)$ which satisfies the mild condition (2.31)[11]. A direct evaluation of the anomaly without using (2.27) is of course possible. We briefly sketch the procedure here , since it justifies our analysis based on the effective expressions in (2.24) ( and partly (2.18) also ). For an operator $O(x,y)$ defined on the lattice, one may define \begin{equation} O_{mn}\equiv \sum_{x,y}\phi_{m}^{\ast}(x)O(x,y)\phi_{n}(y), \end{equation} and the trace \begin{eqnarray} Tr O &=& \sum_{n}O_{nn}\nonumber\\ &=&\sum_{n}\sum_{x,y}\phi_{n}^{\ast}(x)O(x,y)\phi_{n}(y)\nonumber\\ &=&\sum_{x}(\sum_{n,y}\phi_{n}^{\ast}(x)O(x,y)\phi_{n}(y)). \end{eqnarray} The local version of the trace (or anomaly) is then defined by $tr O(x,x) \equiv\\ \sum_{n,y}\phi_{n}^{\ast}(x)O(x,y)\phi_{n}(y)$. For the operator of our interest, we have \begin{equation} tr (-\frac{1}{2}\gamma_{5}D_{W}\frac{1}{\sqrt{D_{W}^{\dagger}D_{W}}})(x) = -\frac{1}{2}\sum_{n=0}^{15} tr \int^{\frac{\pi}{2a}}_{-\frac{\pi}{2a}}\frac{d^{4}k}{(2\pi)^{4}}e^{-ikx}{\cal T}^{-1}_{n}\gamma_{5}D_{W}\frac{1}{\sqrt{D_{W}^{\dagger}D_{W}}}{\cal T}_{n}e^{ikx} \end{equation} where we used the plane wave basis defined in (2.16) combined with the operation ${\cal T}_{n}$. We also used a short hand notation $Oe^{ikx}= \sum_{y}O(x,y)e^{iky}$. We first take the $a\rightarrow 0$ limit of this expression with all $M_{n}, n=0\sim 15$, kept fixed and then take the limit $|M_{n}| \rightarrow \infty$ later. For fixed $M_{n}$ ( to be precise, for fixed $m_{0}/a$ and $r/a$ ), one can confirm that the above integral (2.34) for the domain $\frac{\pi}{2a}\epsilon \leq |k_{\mu}| \leq \frac{\pi}{2a}$ vanishes ( at least ) linearly in $a$ for $a \rightarrow 0$, if one takes into account the trace with $\gamma_{5}$. See also Refs.[7][9]. In the remaining integral \begin{equation} -\frac{1}{2}\sum_{n=0}^{15} (-1)^{n}tr \int^{\frac{\pi}{2a}\epsilon}_{-\frac{\pi}{2a}\epsilon}\frac{d^{4}k}{(2\pi)^{4}}e^{-ikx}\gamma_{5}{\cal T}^{-1}_{n}D_{W}\frac{1}{\sqrt{D_{W}^{\dagger}D_{W}}}{\cal T}_{n}e^{ikx} \end{equation} one may take the limit $a\rightarrow 0$ ( and $\frac{\pi}{2a}\epsilon \rightarrow \infty$ ) with letting $\epsilon$ arbitrarily small. By taking (2.20) into account, one thus recovers the expression (2.29). One can arrive at the same conclusion by using an auxiliary regulator $h(\not \!\! C^{2}/m^{2})$ in the integrand in (2.34) to make the intermediate steps better defined[8]. The domain in (2.17) with arbitrarily small but finite $\epsilon$ thus correctly describes the topological aspects of the continuum limit in the present prescription. Here we went through the details of the anomaly calculation to show that the interpretation of the $N_{\pm}$ states in (A.8) as topological excitations related to species doublers, as is shown in (2.26), is also consistent with the local anomaly calculation. As for a general analysis of chiral anomaly in the overlap operator, see Ref.[15]. At this stage it is instructive to consider an operator defined by \begin{equation} D\equiv \frac{1}{a}[ 1 + (i\not \!\! D + M_{0})\frac{1}{\sqrt{\not \!\! D^{2} + M_{0}^{2}}}] \end{equation} instead of $D$ in (2.24). This $D$ is regarded as an $M_{n}\rightarrow\infty, \ n\neq 0$, limit of the effective operator $D$ (2.24) in the Lagrangian level, and it satisfies ( a continuum version of ) the Ginsparg-Wilson relation (1.1) without any species doubler. The relation $Tr \gamma_{5}=0$ is thus expected to be inconsistent. In fact we have an index related to the chiral Jacobian [5] \begin{eqnarray} Tr \gamma_{5}( 1 - \frac{a}{2}D) &=& \lim_{L\rightarrow large}\sum_{l=1}^{L}\phi_{l}^{\dagger}\gamma_{5}( 1 - \frac{a}{2}D)\phi_{l}\nonumber\\ &=& \sum_{\lambda_{l}=0}\phi_{l}^{\dagger}\gamma_{5}\phi_{l} = n_{+} - n_{-} \end{eqnarray} by noting $\phi_{l}^{\dagger}\gamma_{5}\phi_{l}=0$ for $\lambda_{l}\neq 0$ in $\not \!\! D\phi_{l}=\lambda_{l}\phi_{l}$. On the other hand, if one incorrectly uses $Tr \gamma_{5}=0$ one obtains \begin{eqnarray} Tr \gamma_{5}( 1 - \frac{a}{2}D) &=& -\frac{1}{2}Tr ( \gamma_{5}(i\not \!\! D + M_{0})\frac{1}{\sqrt{\not \!\! D^{2} + M_{0}^{2}}})\nonumber\\ &=& -\frac{1}{2}\lim_{L\rightarrow large}\sum_{l=1}^{L}\phi^{\dagger}_{l} \gamma_{5}M_{0}\frac{1}{\sqrt{\not \!\! D^{2} + M_{0}^{2}}}\phi_{l}\nonumber\\ & =& \frac{1}{2}(n_{+} - n_{-}) \end{eqnarray} by noting $\gamma_{5}\not \!\! D + \not \!\! D\gamma_{5} =0$, $\phi_{l}^{\dagger}\gamma_{5}\phi_{l}=0$ for $\lambda_{l}\neq 0$, and $M_{0}<0$. One thus looses half of the index or anomaly. In this example, the evaluation of $Tr \gamma_{5}$ is somewhat subtle, but $Tr \gamma_{5}=0$ is definitely inconsistent since the calculation in the last line in (2.38) is well-defined. In fact the relations \begin{equation} Tr \gamma_{5} = n_{+} - n_{-}\ \ and \ \ Tr (-\frac{a}{2}\gamma_{5}D) = 0 \end{equation} are consistent for the present operator $D$, since the species doublers at $\gamma_{5}D\phi_{l} = \pm (2/a)\phi_{l}$ are missing. A more rigorously regularized Jacobian for the present example is given by the formula (3.3) to be discussed later. \subsection{General lattice Dirac operator and $Tr \gamma_{5}=0$} We expect that our analysis of $Tr \gamma_{5}=0$, namely its consistency is ensured only by the presence of the would-be species doublers in the Hilbert space, works for a general lattice Dirac operator, since any lattice operator contains $\not \!\! C$ as an essential part. For the smooth near continuum configurations, the lowest dimensional operator $\not \!\! C$ is expected to specify the topological properties. From this viewpoint, the overlap Dirac operator $D$ describes the topological properties such as the index theorem and $Tr \gamma_{5}=0$ in a neater way than the Wilson operator $D_{W}$, mainly because the operator $D$ projects all the species doublers to the vicinity of $2/a$: The behavior for small values of $\not \!\! C$ (i.e.,for $|\not \!\! C| \ll 1/a$ ) is described in a more clear-cut way by $D$, and one can recognize clearly the topological $N_{\pm}$ states related to species doublers. We here note that the Pauli-Villars regularization in continuum theory can be analyzed in a similar way. The Pauli-Villars regulator is defined in the path integral by introducing a bosonic spinor $\phi$ into the action \begin{equation} S = \int d^{4}x [\bar{\psi}(i\not \!\! D - m )\psi + \bar{\phi}(i\not \!\! D - M ) \phi ]. \end{equation} The Jacobian for the global chiral transformation then gives rise to the graded trace[11] \begin{equation} Tr \gamma_{5} = Tr_{\psi}\gamma_{5} - Tr_{\phi}\gamma_{5} = 0. \end{equation} The relation $Tr \gamma_{5} = 0$ is thus consistent with any topologically non-trivial background gauge field because of the presence of the unphysical regulator $\phi$. This $\phi$ is analogous to the species doublers in lattice regularization. \section{Implications of the present analysis} We have shown that the consistency of $Tr \gamma_{5} = 0$ for topologically non-trivial background gauge fields requires the presence of some unphysical states in the Hilbert space. Coming back to the original lattice theory defined by \begin{equation} S = \bar{\psi}D\psi \end{equation} with $D$ satisfying the relation (1.1), one obtains twice of (1.3) as a Jacobian factor for the global chiral transformation [5] $\delta\psi = i\epsilon\gamma_{5}(1-\frac{a}{2}D)\psi$ and $\delta\bar{\psi} = \bar{\psi}i\epsilon (1-\frac{a}{2}D)\gamma_{5}$, which leaves the action (3.1) invariant. One can rewrite (1.3) as \begin{equation} Tr \gamma_{5}( 1 - \frac{a}{2}D) = \tilde{T}r \gamma_{5}( 1 - \frac{a}{2}D) = \tilde{T}r \gamma_{5} = n_{+} - n_{-} \end{equation} where the modified trace $\tilde{T}r$ is defined by truncating the unphysical $N_{\pm}$ states with $\lambda_{n} =\pm 2/a$ . Without the $N_{\pm}$ states, $\tilde{T}r \gamma_{5}\frac{a}{2}D = 0$ since the eigenvalues $\lambda_{n}$ of $\gamma_{5}D$ with $\lambda_{n} \neq 0, \pm 2/a$ appear always pairwise at $\pm |\lambda_{n}|$. See Appendix. If one takes a smooth continuum limit of $\tilde{T}r \gamma_{5}= n_{+} - n_{-}$ in (3.2) , one recovers the result of the continuum path integral (1.6). If one considers that $\tilde{T}r \gamma_{5}$ is too abstract, one may define it more concretely by \begin{eqnarray} Tr \gamma_{5}( 1 - \frac{a}{2}D) f(\frac{(\gamma_{5}D)^{2}}{M^{2}})&=& \tilde{T}r \gamma_{5}( 1 - \frac{a}{2}D)f(\frac{(\gamma_{5}D)^{2}}{M^{2}})\nonumber\\ &=& \tilde{T}r \gamma_{5}f(\frac{(\gamma_{5}D)^{2}}{M^{2}}) = n_{+} - n_{-} \end{eqnarray} for {\em any } $f(x)$ which satisfies the mild condition in (2.31). See also (1.3) and (1.5). This relation suggests that we can extract the local index ( or anomaly) by \begin{equation} tr \gamma_{5}( 1 - \frac{a}{2}D) f(\frac{(\gamma_{5}D)^{2}}{M^{2}})(x) \end{equation} which is shown to be independent of the choice of $f(x)$ in the limit $a \rightarrow 0$ and leads to (2.30) ( for $f(x)$ which goes to zero rapidly for $x\rightarrow \infty$ ) by using only the general properties of $D$ [8]. If one constrains the momentum domain to (2.16) from the beginning , one may use the last expression in (3.3) to evaluate the anomaly for a more general class of $f(x)$. We thus naturally recover the result of the continuum path integral[11]. As for a more practical implication of our analysis of $Tr \gamma_{5} = 0$ in lattice theory, one may say that any result which depends critically on the states $N_{\pm}$ is {\em unphysical}. It is thus necessary to define the scalar density ( or mass term ) and pseudo-scalar density in the theory (3.1) by [16][6] \begin{eqnarray} S(x) &=& \bar{\psi}_{L}\psi_{R} + \bar{\psi}_{R}\psi_{L}= \bar{\psi}( 1 - \frac{a}{2}D)\psi\nonumber\\ P(x) &=& \bar{\psi}_{L}\psi_{R} - \bar{\psi}_{R}\psi_{L}= \bar{\psi}\gamma_{5}( 1 - \frac{a}{2}D)\psi \end{eqnarray} Here we defined two independent projection operators \begin{eqnarray} P_{\pm} &=& \frac{1}{2}( 1 \pm \gamma_{5})\nonumber\\ \hat{P}_{\pm} &=& \frac{1}{2}( 1 \pm \hat{\gamma}_{5}) \end{eqnarray} with $\hat{\gamma}_{5} = \gamma_{5}(1 - aD)$ which satisfies $\hat{\gamma}_{5}^{2} = 1$ [6]. The left- and right- components are then defined by \begin{equation} \bar{\psi}_{L,R}= \bar{\psi}P_{\pm}, \ \ \psi_{R,L}= \hat{P}_{\pm}\psi \end{equation} which is based on the decomposition \begin{equation} D = P_{+}D\hat{P}_{-} + P_{-} D\hat{P}_{+}. \end{equation} The physical operators $S(x)$ and $P(x)$ in (3.5) do not contain the contribution from the unphysical states $N_{\pm}$ in (A.8). In the spirit of this construction, the definition of the index by (3.3) which is independent of unphysical states $N_{\pm}$ is natural. In particular, all the unphysical species doublers ( not only the topological ones at $2/a$ ) decouple from the anomaly defined by (3.4) in the limit $a \rightarrow 0$ with fixed $M$. The customary calculation of the index ( and also anomaly ) by the relation[4]-[7][9] \begin{equation} Tr \gamma_{5}( 1 - \frac{a}{2}D) = Tr( - \frac{a}{2}\gamma_{5}D) = n_{+} - n_{-} \end{equation} by itself is of course consistent, since one simply includes the unphysical states $N_{\pm}$ in evaluating $Tr \gamma_{5} = 0$, and consequently one obtains the index $Tr ( - \frac{a}{2}\gamma_{5}D)$ from the unphysical states $N_{\pm}$ only. We after all know that the left-hand side of (3.9) is independent of $N_{\pm}$. Rather, the major message of our analysis is that the continuum limit of $Tr \gamma_{5} = 0$ in (1.8) ( unlike the relation $\tilde{T}r \gamma_{5} = n_{+} - n_{-}$ ) {\em cannot } be defined in a consistent way when the (would-be) species doublers disappear from the Hilbert space. It is clear from the expression of $Tr \gamma_{5} = 0$ in (1.8) that the $a\rightarrow 0$ limit of $Tr \gamma_{5} = 0$ is not defined consistently. One may then ask how the calculation of local anomaly on the basis of (3.9) could be consistent in the limit $a\rightarrow 0$ if $Tr \gamma_{5} = 0$ is inconsistent? A key to resolve this apparent paradox is the failure of the decoupling of heavy fermions in the evaluation of anomaly. The massive unphysical species doublers do not decouple from the anomaly , as is seen in (2.29), for example. If one insists on $Tr \gamma_{5} = 0$ in the continuum limit, one is also insisting on the failure of the decoupling of these infinitely massive particles from $Tr \gamma_{5} = 0$. The contributions of these heavy fermions to the anomaly! and to $Tr \gamma_{5} = 0$ precisely cancel, just as in the case of the evaluation of global index in (3.9). Namely, the local anomaly itself is {\em independent} of these massive species doublers in the continuum limit, as is clear in (3.4). In this sense, (3.4) is the only logically consistent definition of local anomaly. It is an advantage of the finite lattice formulation that we can now clearly illustrate this subtle cancellation of the contributions of those ultra-heavy regulators to $Tr \gamma_{5} = 0$ and anomaly on the basis of (1.8). ( In the case of the Wilson fermion operator $D_{W}$, an analogous cancellation takes place in $Tr \gamma_{5} + pseudo-scalar\ \ mass\ \ term$ induced by the chiral variation of the action.) When one defines a chiral theory by recalling (3.8) [6][17] \begin{equation} S = \bar{\psi} P_{+}D\hat{P}_{-}\psi = \bar{\psi}_{L} D \psi_{L} \end{equation} one obtains the {\em covariant} gauge anomaly (or Jacobian) \begin{equation} tr T^{a}\gamma_{5}( 1 - \frac{a}{2}D) = \sum_{n}\phi_{n}(x)^{\dagger}T^{a}\gamma_{5}( 1 - \frac{a}{2}D)\phi_{n}(x) \end{equation} for the gauge transformation $\delta\psi_{L}(x) = i\alpha^{a}(x)T^{a}\hat{P}_{-}\psi_{L}$ and $\delta\bar{\psi}_{L}(x) = \bar{\psi}_{L}P_{+}(-i)\alpha^{a}(x)T^{a}$. An analogue of the $U(1)$ anomaly (3.4) is then defined for the gauge anomaly (3.11) by ( by using $\phi_{n}$ in (1.4)) \begin{equation} \sum_{n}\phi_{n}(x)^{\dagger}T^{a}\gamma_{5}( 1 - \frac{a}{2}D)f(\frac{(\gamma_{5}D)^{2}}{M^{2}})\phi_{n}(x) = \sum_{n}f(\frac{\lambda_{n}^{2}}{M^{2}})\phi_{n}(x)^{\dagger}T^{a}(\gamma_{5} - \frac{a}{2}\lambda_{n})\phi_{n}(x) \end{equation} which reduces to the lattice expression for $M \rightarrow \infty$ with $f(0)=1$. In practice, one first takes the continuum limit $ a \rightarrow 0$ with $M$ fixed and one obtains (see , for example, [8]) \begin{equation} tr T^{a}\gamma_{5}f(\frac{\not \!\! D^{2}}{M^{2}}) \end{equation} which is again known to be independent of the specific choice of $f(x)$ in the limit $M \rightarrow \infty$[11]. For the overlap Dirac operator, one can show that the anomaly calculation in (3.11) by using $tr T^{a}\gamma_{5}=0$ corresponds effectively to a specific choice of $f(x) = 1/\sqrt{ 1 + x}$ in (3.13), just as in the case of $U(1)$ anomaly in (2.30). The definition of the regularized Jacobian (3.12) may be regarded to correspond to the truncation of the states $N_{\pm}$ from the chiral action \begin{equation} S = \sum_{n\in N_{+}} (\frac{2}{a})\bar{C}_{n}C_{n} + \sum_{0\leq\lambda_{n}< 2/a} \lambda_{n}\bar{C}_{n}C_{n} \end{equation} to \begin{equation} \tilde{S} = \sum_{0\leq\lambda_{n}< 2/a}\lambda_{n}\bar{C}_{n}C_{n} \end{equation} and then taking the continuum limit $ a \rightarrow 0$, which is logically more natural as the $N_{\pm}$ states are eliminated from the Hilbert space {\em before} taking the continuum limit. Incidentally, the action (3.14) is obtained from (3.10) by expanding \begin{eqnarray} \bar{\psi}P_{+} &=& \sum _{n} \bar{C}_{n}\bar{v}_{n},\nonumber\\ \hat{P}_{-}\psi &=& \sum _{n} C_{n} v_{n} \end{eqnarray} with the choice of the basis set \begin{eqnarray} \{ v_{j}\} &=& \{ \phi_{n}| \gamma_{5}D\phi_{n}= 0, \gamma_{5}\phi_{n}= - \phi_{n}\}\nonumber\\ &\oplus& \{ \phi_{n}| \gamma_{5}D\phi_{n}= 2/a\phi_{n}, \gamma_{5}\phi_{n}= + \phi_{n}\}\nonumber\\ &\oplus& \{ \hat{P}_{-}\phi_{n}/\sqrt{(1+a\lambda_{n}/2)/2}\ | \gamma_{5}D\phi_{n}= \lambda_{n}\phi_{n}, 2/a > \lambda_{n} > 0 \} \end{eqnarray} \begin{eqnarray} \{ \bar{v}_{k}^{\dagger}\} &=& \{ \phi_{n}| \gamma_{5}D\phi_{n}= 0, \gamma_{5}\phi_{n}= + \phi_{n}\}\nonumber\\ & \oplus& \{ \phi_{n}| \gamma_{5}D\phi_{n}= 2/a\phi_{n}, \gamma_{5}\phi_{n}= + \phi_{n}\}\nonumber\\ &\oplus& \{ P_{+}\phi_{n}/\sqrt{(1+a\lambda_{n}/2)/2}\ | \gamma_{5}D\phi_{n}= \lambda_{n}\phi_{n}, 2/a > \lambda_{n} > 0 \} \end{eqnarray} in terms of the eigenstates of $\gamma_{5}D\phi_{n} = \lambda_{n}\phi_{n}$ summarized in Appendix. Consequently, the path integral for a fixed background gauge field is defined by \begin{eqnarray} Z &=& J \int \prod_{n\in N_{+}}d\bar{C}_{n}dC_{n}\prod_{0\leq\lambda_{n}< 2/a} d\bar{C}_{n}\prod_{0\leq\lambda_{m}< 2/a}dC_{m}\exp S, \nonumber\\ \tilde{Z} &=& \tilde{J}\int \prod_{0\leq\lambda_{n}< 2/a} d\bar{C}_{n}\prod_{0\leq\lambda_{m}< 2/a}dC_{m}\exp \tilde{S} \end{eqnarray} with Jacobian factors $J$ and $\tilde{J}$ which depend on the basis set. A ( naive ) continuum limit of the truncated expression $\tilde{Z}$ naturally gives rise to the covariant path integral formulation of chiral gauge theory[11]. In particular, the fermion number anomaly which is given by (3.3) gives rise to the fermion number violation in chiral gauge thoery. As is well-known, this formulation of the continuum limit is consistent if the anomaly cancellation condition $tr T^{a}\{ T^{b},T^{c}\} = 0$ is satisfied, when combined with the argument of the robustness of lattice gauge symmetry [18]. See also [19]. An interesting analysis of the definition of chiral theory at a finite $a$ has been given by L\"{u}scher recently [20]. The fermion number violation arises from the non-trivial index of the rectangular ( {\em not} square) matrix in (3.10) (Cf. (2.3)) \begin{equation} dim\ \ ker\ \ \hat{P}_{-}\gamma_{5}DP_{+} - dim\ \ ker\ \ P_{+}\gamma_{5}D\hat{P}_{-} = n_{+} - n_{-} \end{equation} as is seen in the explicit construction of the basis vectors in (3.17) and (3.18): For a general $n\times m$ matrix $M$, one can prove an index theorem \begin{equation} dim\ ker\ M - dim\ ker\ M^{\dagger} = m - n \end{equation} which is a generalization of the case of a square matrix with $m=n$. For the operator $ \hat{P}_{-}\gamma_{5}DP_{+}$ in (3.20), the dimensions of the column and row vectors are respectively given by using the projection operators as $Tr \hat{P}_{-}$ and $Tr P_{+}$, and thus $ m - n = Tr P_{+} - Tr \hat{P}_{-} = Tr \gamma_{5}(1-\frac{a}{2}D) = n_{+} - n_{-}$ [20]. Incidentally, an analogous analysis provides an alternative proof of the equivalence of $Tr\gamma_{5}=0$ with the index relation (2.3). \section{Discussion and conclusion} Motivated by the recent interesting developments in lattice gauge theory, we analyzed the physical implications of the condition $Tr \gamma_{5} = 0$ in detail. We have shown that $Tr \gamma_{5} = 0$, whose validity is often taken for granted, is consistent only when one includes some unphysical states in the Hilbert space. The continuum $a\rightarrow 0$ limit of $Tr \gamma_{5} = 0$ is not defined consistently as is seen in (1.8). We have explained that the failure of the decoupling of heavy fermions in the anomaly calculation is a key to understand the consistency of the customary lattice calculation of anomaly where $Tr \gamma_{5} = 0$ is used. Our analysis is perfectly consistent with the relation (1.6) in the continuum path integral and even provides positive support for the formula (3.3) and the related definition of anomaly (3.4) in lattice theory. We here want to comment on an analysis of the photon phase operator[21] where a closely related phenomenon associated with the notion of index takes place [22]. The Maxwell field is expanded into an infinite set of harmonic oscillators, and thus the analysis of the photon phase operator is performed for a simple harmonic oscillator \begin{equation} H = \frac{1}{2}( p^{2} + \omega^{2}q^{2}) = \hbar \omega ( a^{\dagger}a + \frac{1}{2}) \end{equation} The quantum requirement of the absence of the negative normed states leads to $a|0\rangle = 0$, and thus the index relation \begin{equation} dim \ \ ker \ \ a - dim \ \ ker \ \ a^{\dagger} = 1 \end{equation} since no states are annihilated by $ a^{\dagger}$. On the other hand, the existence of the observable {\em hermitian} phase operator $\varphi$ requires a decomposition [21] \begin{equation} a = U(\varphi)\sqrt{N}, \ \ \ a^{\dagger} = \sqrt{N}U(\varphi)^{-1} \end{equation} with a {\em unitary} $U(\varphi) = e^{i\varphi}$ and $N = a^{\dagger}a$. These expressions suggest \begin{equation} dim \ \ ker \ \ a - dim \ \ ker \ \ a^{\dagger} = 0 \end{equation} in contradiction to the relation (4.2), since the unitary factor $U(\varphi)$ does not influence the analysis of index. The index (4.2) thus provides a no go theorem against the hermitian photon phase operator and the resulting familiar phase-number uncertainty relation [22]. To circumvent the topological stricture (4.2), one may truncate the operator $a$ to an $(s+1)\times (s+1)$ dimensional square matrix \begin{eqnarray} a_{s}&=&\left(\begin{array}{cccccc} 0&1&0 &0 &..&0\\ 0&0 &\sqrt{2}&0 &..&0\\ 0&0 &0 &\sqrt{3}&..&0\\ .&.&.&. &. &.\\ 0&0 &0 &0&..&\sqrt{s} \\ 0&0 &0 &0&..&0 \end{array}\right)\nonumber\\ &=& |0\rangle\langle 1| + |1\rangle\langle 2|\sqrt{2} .... + |s-1\rangle\langle s|\sqrt{s} \end{eqnarray} and $a_{s}^{\dagger} = (a_{s})^{\dagger}$. One then obtains a vanishing index for a finite dimensional square matrix [22] \begin{equation} dim \ \ ker \ \ a_{s} - dim \ \ ker \ \ a_{s}^{\dagger} = 0 \end{equation} and one can in fact introduce a hermitian phase operator $\phi$ [23] which satisfies the relation $a_{s}= e^{i\phi}\sqrt{a_{s}^{\dagger}a_{s}}$. The parameter $s$ or the state $|s\rangle$ stands for the cut-off parameter analogous to the $N_{\pm}$ states related to $Tr \gamma_{5} = 0$ in lattice theory. A careful analysis of the uncertainty relation shows that the hermitian operator $\phi$ , when used to analyze the data which is already in the quantum limit, leads to a substantial deviation from the minimum uncertainty relation at the characteristically quantum domain with {\em small} average photon numbers. This artificial deviation from the minimum uncertainty is caused by the presence of the unphysical cut-off introduced by $|s\rangle$, which fails to decouple from the low energy quantities for arbitrarily large but finite $s$ [22]. Also, a large $s$ limit of (4.6) is not defined consistently, which is analogous to the ill-defined continuum limit of $Tr \gamma_{5}= 0$ in (1.8). It is expected that an analogous unphysical result will appear in lattice gauge theory if one analyzes the low energy quantity which critically depends on the unphysical states $N_{\pm}$. In fact , it is known that one {\em has to } eliminate the contribution of the $N_{\pm}$ states to the physical observables such as $S(x)$ and $P(x)$ in (3.5) [16][6].
\section*{I. Introduction} Numerical simulations of the QCD (Quantum Chromodynamics) equation of state on the lattice predict that at very high density and/or temperature hadronic matter undergoes a phase transition to Quark Gluon Plasma (QGP)~\cite{ukawa}(see Fig.~\ref{QCDfig}). One expects that ultrarelativistic heavy ion collisions might create conditions conducive for the formation and study of QGP. Various model calculations have been performed to look for observable signatures of this state of matter. However, among various signatures of QGP, photons and dileptons are known to be advantageous, primarily so because, electromagnetic interaction could lead to detectable signal. However, it is weak enough to let the produced particles (real photons and dileptons) escape the system without further interaction and thus carrying the information of the constituents and their momentum distribution in the thermal bath. The disadvantage with photons is the substantial background from various processes (thermal and non-thermal)~\cite{janepr}. Among these, the contribution from hard QCD processes is well understood in the framework of perturbative QCD and the yield from hadronic decays e. g. $\pi^0\,\rightarrow\,\gamma\,\gamma$ can be accounted for by invariant mass analysis. However, photons from the thermalised hadronic gas pose a more difficult task to disentangle. Therefore it is very important to estimate photons from hot and dense hadronic gas along with the possible modification of the hadronic properties. \begin{figure} \centerline{\psfig{figure=qcd.ps,height=6cm,width=8cm}} \caption{ QCD phase diagram.} \label{QCDfig} \end{figure} The importance of the electromagnetic probes for the thermodynamic state of the evolving matter was first proposed by Feinberg in 1976~\cite{feinberg}. While for most purposes one can calculate the emission rates in a classical framework, Feinberg showed that the emission rates can be related to the electromagnetic current current correlation functions in a thermalised system in a quantum picture and, more importantly, in a nonperturbative manner. Generally the production of a particle which interact weakly with the constituents of the thermal bath ( the constituents may interact strongly among themselves, we need not give any explicit form of their coupling strength) can always be expressed in terms of the discontinuities or imaginary parts of the self energies of that particle~\cite{ruuskanen,bellac} which never thermalizes. We should, therefore look at the connection between the electromagnetic emission rates (real photons and lepton pairs, which correspond to virtual photons) and the photon spectral function ( which is connected with the discontinuities in self energies) in a thermal system~\cite{weldon90}, which in turn connected to the hadronic electromagnetic current current correlation tensor~\cite{mclerran} through Maxwell equations. We will show below that the photon emission rate can be obtained from the dilepton (which originate from a virtual photon) emission rate with little effort. In section II we present the general formalism for the dilepton and photon production rate from a thermal bath. In section III we calculate the medium modifications of hadrons. Section IV is devoted to discuss the results of our calculations. In section V we present a summary and discussion. \section*{II. Electromagnetic Probes - Formulation} We begin our discussion with the dilepton production rate. In the following we will follow the work of Weldon~\cite{weldon90} for the derivation of dilepton emission rate. Let $e_0\,J_\mu\,A^\mu$ be the interaction between the photon and the particles in the heat bath, then to lowest order in the electromagnetic coupling, the $S$ matrix element for the transition $\mid I\rangle\,\rightarrow\,\mid H;l^+l^-\rangle$ is given by \begin{equation} S_{HI}=e_0\,\int\,\langle\,H;l^+l^-\,\mid\,J_\mu\,A^\mu(x)\,\mid\,I\rangle\, d^4x\,e^{iq\cdot x} \end{equation} where $\mid I\rangle$ is the initial state corresponding to the two incoming nuclei, $\mid H;l^+l^-\rangle$ is the final state which corresponds to a lepton pair plus anything (Hadronic), the parameter $e_0$ is the bare (unrenormalised) charge and $q=p_1\,+\,p_2$ is the four momentum of the lepton pair. Since we assume that the lepton pair do not interact with the emitting system, the matrix element can be factorised in the following way, \begin{equation} \langle\,H;l^+l^-\mid\,J_\mu\,A^\mu(x)\,\mid\,I\rangle= \langle H\mid\,A^\mu(x)\mid\,I\rangle\langle\,l^+l^-\mid\,J_\mu\,\mid\,0\rangle \end{equation} Putting the explicit form of the current $J^\mu$ in terms of the Dirac spinors we obtain \begin{equation} S_{HI}=e_0\,\frac{\bar{u}(p_1)\gamma_\mu\,v(p_2)}{V\sqrt{2E_12E_2}}\int d^4x\,e^{iq\cdot x}\langle H\mid\,A^\mu(x)\mid I\rangle \end{equation} Therefore dilepton multiplicity from a thermal system is obtained as, (by summing over the final states and averaging over the initial states with a weight factor $Z(\beta)^{-1}\,e^{-\beta\,E_I}$) \begin{equation} N=\frac{1}{Z(\beta)}\sum_I\,\sum_H\,\mid S_{HI}\,\mid^2\,e^{-\beta\,E_I} \end{equation} where $E_I$ is the total energy in the initial state. After some algebra $N$ can be written in a compact form as follows, \begin{equation} N=e_0^2\,L^{\mu\nu}\,H_{\mu\nu} \frac{d^3p_1}{(2\pi)^32E_1} \,\frac{d^3p_2}{(2\pi)^32E_2} \end{equation} where $L_{\mu\nu}$ is the leptonic tensor given by \begin{eqnarray} L^{\mu\nu}&=&\frac{1}{4}\, \sum_{spins}\,\bar{u}(p_1)\gamma^\mu\,v(p_2)\bar{v}(p_2) \gamma^\nu\,u(p_1)\nonumber\\ &=&\,p_1^\mu\,p_2^\nu+p_2^\mu\,p_2^\nu-\frac{q^2}{2}\,g^{\mu \nu} \end{eqnarray} and $H_{\mu\nu}$ is the photon tensor \begin{equation} H_{\mu \nu}=\frac{1}{Z(\beta)}\,e^{-\beta q_0}\sum_H\,\int d^4x\,d^4y\, e^{iq\cdot (x-y)} \,\langle\,H\mid\,A_\mu(x)\,A_\nu(y)\mid\,H\rangle\,e^{-\beta\,E_H} \end{equation} to obtain the above equation we have used the resolution of identity $1=\sum_I\,\mid I\rangle\langle I\mid$, and the energy conservation equation $E_I=E_H+q_0$ where $q_0$ is the energy of the lepton pair and $E_H$ is the energy of the rest of the system produced after collision. Using the translational invariance of the matrix element we can write \begin{equation} H_{\mu\nu}=2\pi\,\Omega\,e^{-\beta\,q_0}\rho^{\mu\nu} \end{equation} where $\Omega$ is the four volume of the system and $\rho^{\mu\nu}$ is the spectral function of the photon in the heat bath, \begin{equation} \rho_{\mu \nu}(\vec q,q_0)=\frac{1}{2\pi\,Z(\beta)}\int d^4x\,e^{iq\cdot x}\sum_H \,\langle\,H\mid\,A_\mu(x)\,A_\nu(0)\mid\,H\rangle\,e^{-\beta\,E_H} \label{spectral1} \end{equation} The rate of dilepton production per unit volume ($N/\Omega$) is given by, \begin{equation} \frac{dN}{d^4x}=2\pi\,e_0^2 L_{\mu \nu}\,e^{-\beta q_0}\rho^{\mu \nu}(q) \frac{d^3p_1}{(2\pi)^32E_1}\,\frac{d^3p_2}{(2\pi)^32E_2} \label{funda} \end{equation} This result which expresses the dilepton emission rate in terms of the spectral function of the photon in the medium is a fundamental result. At zero temperature ($\beta\,\rightarrow\,\infty$) the only state which enters in the spectral function ($\rho(s)$) is the vacuum and the spectral function can be expressed in terms of the imaginary part of Greens function (i.e the discontinuity along the real axis) $\rho(s)=-\frac{1}{\pi}\,{\rm Im}\,G(s)$~\cite{lsbrown}. Unlike at zero temperature, in the case of non-zero temperature the time ordered ($D^{\mu\nu}$) and the Feynman propagator ($D^{\mu \nu}_F$) are different. The relation between them is given by~\cite{nieves}, \begin{equation} {\rm Im}\,D^{\mu \nu}=(1+2f_{BE}){\rm Im}\,D^{\mu \nu}_F \label{toptof} \end{equation} where $f_{BE}$ is the thermal distribution for photons. Now the time ordered Green's function can be expressed in terms of the spectral function as follows \begin{equation} D^{\mu \nu}(q_0,\vec q)=\int_{-\infty}^{\infty}\,d\alpha\, \left[\frac{\rho^{\mu \nu}(\alpha,\vec q)}{q^0-\alpha+i\epsilon} -\frac{\rho^{\nu\mu}(\alpha,-\vec q)}{q^0+\alpha-i\epsilon}\right] \label{top} \end{equation} Using the KMS relation one can show that \begin{equation} \rho^{\mu \nu}(q_0,\vec q)=e^{\beta\,q_0}\rho^{\nu\mu}(-q_0,-\vec q) \label{kms} \end{equation} substituting Eq.~(\ref{kms}) in Eq.~(\ref{top}) and using Eq.~(\ref{toptof}) we obtain \begin{equation} \rho^{\mu \nu}(q_0,\vec q)= \frac{-1}{\pi}\,(1+f_{BE}(q_0)) {\rm Im}\,D^{\mu \nu}_F(q_0,\vec q) \label{spectral} \end{equation} The above equation implies that to evaluate the spectral function at $T\neq 0$ we need to know the imaginary part of the Feynman propagator. It is important to note that above expression for spectral function reduces its vacuum value as $\beta\rightarrow\,\infty$. The Feynman propagator can be expressed in terms of the proper self energy as \begin{equation} D^{\mu \nu}_F=-\left[\,\frac{Z_3\,A^{\mu \nu}}{q^2-\Pi_T}+ \frac{Z_3\,B^{\mu \nu}}{q^2-\Pi_L}\right] + (\zeta-1)\frac{q_\mu\,q_\nu}{q^4} \label{selfe} \end{equation} where $Z_3$ is the wave function renormalization constant at zero temperature, $A^{\mu \nu}$ and $B^{\mu \nu}$ are transverse and longitudinal projection tensors respectively. The presence of the parameter $\zeta$ indicates the gauge dependence of the propagator. Although the gauge dependence cancels out in the calculation of physical quantities, one should, however, be careful when extracting physical quantities from the propagator directly, especially in the non-abelian gauge theory. Using Eqs.~(\ref{funda}), (\ref{spectral}) and (\ref{selfe}) we get \begin{equation} \frac{dN}{d^4x}=2\,e^2\,L_{\mu \nu}(A^{\mu \nu}\rho_T\,+\,B^{\mu \nu}\,\rho_L) \frac{d^3p_1}{(2\pi)^32E_1}\,\frac{d^3p_2}{(2\pi)^32E_2}\, f_{BE}(q_0) \label{dilrate} \end{equation} where $e^2=Z_3\,e_0^2$, is the physical charge of the electron and \begin{equation} \rho_k=\frac{{\rm Im}\,\Pi_k}{(q^2-Re\,\Pi)^2+{\rm Im}\,\Pi_k^2} \end{equation} The emission rate of dilepton can also be expressed in terms of the electromagnetic current-current correlation functions~\cite{mclerran}. Denoting the hadronic part of the electromagnetic current operator by $J^h_\mu$, the leptonic part by $J^l_\nu$ and the free photon propagator by $D^{\mu \nu}$, we have the matrix element for this transition : \begin{equation} S_{HI}=\langle\,H;l^+\,l^-\,\mid\int d^4x d^4y J^h_{\mu}(x)D^{\mu \nu}\,J^l_{\nu}(y)\mid I\rangle \end{equation} As in the earlier case the leptonic part of the current can be easily factored out. Writing the Fourier transform of photon propagator and squaring the matrix elements, one obtains, for the rate of dilepton production, \begin{equation} dR=e_0^2\,L^{\mu \nu}\,W_{\mu \nu}\,\frac{1}{q^4}\, \frac{d^3\,p_1}{(2\pi)^3\,E_1} \frac{d^3\,p_2}{(2\pi)^3\,E_2} \label{dnd4x} \end{equation} where $W^{\mu\nu}(q)$ is just the Fourier transform of the thermal expectation value of the real time electromagnetic current-current correlation function : \begin{equation} W^{\mu\nu}=\int d^4x e^{-iqx}\sum_H\, \langle H\,\mid J^{\mu}(x)J^{\nu}(0)\mid\,H\,\rangle\, \frac{e^{-\beta\,E_H}}{Z} \label{correl} \end{equation} The subtleties of the thermal averaging have been elucidated earlier. It is thus readily seen (eq.~\ref{dnd4x}) that the dilepton data yields considerable information about the thermal state of the hadronic system; at least the full tensor structure of $W^{\mu\nu}$ can in principle be determined. The most important point to realize here is that the analysis is essentially nonperturbative. At this point one may wonder what is the connection between the electromagnetic current-current correlation function and the spectral function? The connection can be expressed in a straight forward way by substituting $J_\mu$ and $J_\nu$ using Maxwell equation ($\partial_\alpha\partial^\alpha\,A^\mu -\partial^\mu\,(\partial_\alpha\,A^\alpha)=j^\mu$) in Eq.~(\ref{correl}) to obtain, \begin{equation} e_0^2\,W^{\mu \nu}=2\pi\,(q^2\,g^{\mu\alpha}-q^\mu\,q^\alpha)\, \rho_{\alpha\beta}\,(q^2\,g^{\beta\nu}-q^\beta\,q^\nu) \end{equation} In the literature most of the dilepton production rate from a thermal system are calculated with the approximation $\Pi_T=\Pi_L$ which gives the spectral function as \begin{equation} \rho^{\mu \nu}(q_0,\vec q)=(-g^{\mu \nu}+q^\mu\,q^\nu/q^2) \frac{-1}{\pi}\,(1+f_{BE}(q_0))\,{\rm Im}\,\left[\frac{1}{q^2-\Pi_L}\right] \label{restspec} \end{equation} Since ${\rm Im}\,\Pi_k$ and ${\rm Re}\,\Pi_k$ are proportional to $\alpha$ (the fine structure constant) they are very small for most of our practical purposes. Therefore $\rho_k$ is given by \begin{equation} \rho_k=\frac{{\rm Im}\,\Pi_k}{q^4} \label{approx} \end{equation} Using Eqs.~(\ref{dilrate}), (\ref{restspec}), (\ref{approx}) and the following result \begin{eqnarray} \int\,\prod_{i=1,2}\,\frac{d^3p_i}{(2\pi)^32E_i} \delta^4(p_1+p_2-q)\,L^{\mu \nu}(p_1,p_2)&=&\frac{1}{(2\pi)^6}\frac{2\pi}{3} (1+\frac{2m^2}{q^2})\nonumber\\ &&\times\sqrt{1-\frac{4m^2}{q^2}}C^{\mu \nu} \end{eqnarray} where $C^{\mu \nu}=q^\mu\,q^\nu-q^2\,g^{\mu \nu}$, we get, \begin{equation} \frac{dR}{d^4q}=\frac{\alpha}{12\pi^4\,q^2}(1+\frac{2m^2}{q^2}) \sqrt{1-\frac{4m^2}{q^2}}{\rm Im}\Pi^\mu_\mu\,f_{BE}(q_0) \label{usedrate} \end{equation} This is the familiar results most widely used for dilepton emission rate~\cite{bellac}. To obtain real photon emission rate per unit volume ($dR$) from a system under thermal equilibrium we note that the dilepton emission rate differs from the photon emission rate in the following way. The factor $e^2\,\,L_{\mu \nu}\, \frac{1}{q^4}$ is the product of the electromagnetic vertex $\gamma^\ast\,\rightarrow\,l^+\,l^-$, the leptonic current involving Dirac spinors and the square of the photon propagator should be replaced by the factor $\sum\,\epsilon_\mu\,\epsilon_\nu (=-g_{\mu \nu})$ for the real (on-shell) photon. Finally the phase space factor $d^3p_1/[(2\pi)^32\,E_1]\,d^3p_2/[(2\pi)^32E_2]$ should be replaced by, $d^3q/[(2\pi)^32q_0]$ to obtain \begin{equation} dR=-\frac{1}{(2\pi)^3}\,g_{\mu \nu}\,W^{\mu \nu}\,\frac{d^3q}{q_0} \end{equation} The current-current correlation function is related to the photon self energy as~\cite{bellac,beres,kapusta} \begin{equation} W_{\mu \nu}=2\,f_{BE}(q_0)\,{\rm Im}\,\Pi_{\mu \nu} \end{equation} Therefore, \begin{equation} q_0\frac{dR}{d^3q}=-2\frac{1}{(2\pi)^3}\,{\rm Im}\Pi^\mu_\mu\,f_{BE}(q_0) \end{equation} The above equation can also be obtained directly from eq.~\ref{usedrate}. The emission rate given above is correct up to order $e^2$ in electromagnetic interaction but exact, in principle, to all order in strong interaction. But for all practical purposes one is able to evaluate up to a finite order of loop expansion. Now it is clear from the above results that to evaluate photon and dilepton emission rate from a thermal system we need to evaluate the imaginary part of the photon self energy. The Cutkosky rules at finite temperature~\cite{kobes,adas,gelis} gives a systematic procedure to calculate the imaginary part of a Feynman diagram. The Cutkosky rule expresses the imaginary part of the $n$-loop amplitude in terms of physical amplitude of lower order ($n-1$ loop or lower). This is shown schematically in Fig.~\ref{opt}. When the imaginary part of the self energy is calculated up to and including $L$ order loops where $L$ satisfies $x\,+\,y\,<\,L\,+\,1$ then we obtain the photon emission rate for the reaction; $x$ particles $\rightarrow$ $y$ particles $+\,\gamma$ and the above formalism becomes equivalent to the relativistic kinetic theory formalism~\cite{gale}. \begin{figure} \centerline{\psfig{figure=opt.ps,width=15cm,height=6cm,angle=-90}} \caption{Optical Theorem in Quantum Field Theory} \label{opt} \end{figure} For a reaction $1\,+\,2\,\rightarrow\,3\,+\gamma$ the photon (of energy E) emission rate is given by~\cite{sourav}, \begin{eqnarray} E\frac{dR}{d^3p}&=&\frac{N}{16(2\pi)^7E}\,\int_{(m_1+m_2)^2}^{\infty} \,ds\,\int_{t_{\rm {min}}}^{t_{\rm {max}}}\,dt\,|{\cal M}|^2\, \int\,dE_1\nonumber\\ &&\times\int\,dE_2\frac{f(E_1)\,f(E_2)\left[1+f(E_3)\right]}{\sqrt{aE_2^2+ 2bE_2+c}} \label{rktp} \end{eqnarray} where \begin{eqnarray} a&=&-(s+t-m_2^2-m_3^2)^2\nonumber\\ b&=&E_1(s+t-m_2^2-m_3^2)(m_2^2-t)+E[(s+t-m_2^2-m_3^2)(s-m_1^2-m_2^2)\nonumber\\ &&-2m_1^2(m_2^2-t)]\nonumber\\ c&=&-E_1^2(m_2^2-t)^2-2E_1E[2m_2^2(s+t-m_2^2-m_3^2)-(m_2^2-t)(s-m_1^2-m_2^2)] \nonumber\\ &&-E^2[(s-m_1^2-m_2^2)^2-4m_1^2m_2^2]-(s+t-m_2^2-m_3^2)(m_2^2-t)\nonumber\\ &&\times(s-m_1^2-m_2^2) +m_2^2(s+t-m_2^2-m_3^2)^2+m_1^2(m_2^2-t)^2\nonumber\\ E_{1{\rm {min}}}&=&\frac{(s+t-m_2^2-m_3^2)}{4E}+\frac{Em_1^2}{s+t-m_2^2-m_3^2} \nonumber\\ E_{2{\rm {min}}}&=&\frac{Em_2^2}{m_2^2-t}+\frac{m_2^2-t}{4E}\nonumber\\ E_{2{\rm {max}}}&=&-\frac{b}{a}+\frac{\sqrt{b^2-ac}}{a}\nonumber \end{eqnarray} In a similar way the dilepton emission rate for a reaction $a\,\bar a\,\rightarrow\,l^+\,l^-$ can be obtained as \begin{eqnarray} \frac{dN}{d^4xd^3qdM}&=&\int {d^3p_a\over 2E_a(2\pi)^3}f(p_a) \int {d^3p_{\bar a}\over 2E_{\bar a}(2\pi)^3}f(p_{\bar a}) \int {d^3p_1\over 2E_1(2\pi)^3} \int {d^3p_2\over 2E_2(2\pi)^3}\nonumber\\ &&\mid M\mid_{a\bar a\rightarrow l^+l^-}^2 (2\pi)^4\delta(\vec p_a+\vec p_{\bar a}-\vec p_1-\vec p_2) \delta(E_a+E_{\bar a}-E_1-E_2)\nonumber\\ &&\delta(p_a+p_{\bar a}) \delta(M-E_a-E_{\bar a}) \end{eqnarray} where $M^2=(p_a+p_{\bar a})^2$ is the invariant mass and $f(p_a)$ is the occupation probability in the momentum space. The Pauli blocking of the lepton pair in the final state has been neglected in the above equation. \section*{III. Finite Temperature Properties} To calculate the effective mass and decay widths of vector mesons we begin with the following $VNN$ (Vector - Nucleon - Nucleon) Lagrangian density: \begin{equation} {\cal L}_{VNN} = g_{VNN}\,\left({\bar N}\gamma_{\mu} \tau^a N{V}_{a}^{\mu} - \frac{\kappa_V}{2M}{\bar N} \sigma_{\mu \nu}\tau^a N\partial^{\nu}V_{a}^{\mu}\right), \label{lag1} \end{equation} where $V_a^{\mu} = \{\omega^{\mu},{\vec {\rho}}^{\mu}\}$, $M$ is the free nucleon mass, $N$ is the nucleon field and $\tau_a=\{1,{\vec {\tau}}\}$. The values of the coupling constants $g_{VNN}$ and $\kappa_V$ will be specified later. With the above Lagrangian we proceed to calculate the $\rho$-self energy. \begin{equation} \Pi_{\mu \nu} = -2ig_{VNN}^2\,\int\, \frac{d^4{p}}{(2\pi)^4}\,K_{\mu \nu}(p,k), \label{pimunu} \end{equation} with, \begin{equation} K_{\mu \nu} = \frac{{\rm Tr}\left[\Gamma_{\mu}(k)\,(p\!\!\!/+M^{*}) \Gamma_{\nu}(-k)\,(p\!\!\!/-k\!\!\!/+M^{*})\right]}{(p^2-M^{* 2}) ((p-k)^2-M^{*2})}, \label{smn} \end{equation} The vertex $\Gamma_{\mu}(k)$ is calculated by using the Lagrangian of Eq.~(\ref{lag1}) and is given by \begin{equation} \Gamma_{\mu}(k) = \gamma_{\mu} + \frac{i\kappa_V}{2M}\sigma_{\mu \alpha} k^{\alpha}. \label{vertex} \end{equation} Here $M^{\ast}$ is the in-medium (effective) mass of the nucleon at finite temperature which we calculate using the Mean-Field Theory (MFT)~\cite{vol16}. The value of $M^{\ast}$ can be found by solving the following self consistent equation: \begin{equation} M^{\ast} = M -\frac{4g_s^2}{m_s^2}\,\int\,\frac{d^3{\bf p}}{(2\pi)^3}\, \frac{M^{\ast}}{({\bf p}^2+M^{\ast 2})^{1/2}}\,\left[f_N(T)+f_{\bar N}(T)\,\right], \label{MN} \end{equation} where $f_N(T)(f_{\bar N}(T))$ is the Fermi-Dirac distribution for the nucleon (antinucleon), $m_s$ is mass of the neutral scalar meson ($\sigma$) field, and, the nucleon interacts via the exchange of isoscalar meson with coupling constant $g_s$. Since the exact solution of the field equations in QHD is untenable, these are solved in mean field approximation. In a mean field approximation one replaces the field operators by their ground state expectation values which are classical quantities; this renders the field equations exactly solvable. The vacuum part of the $\rho$ self energy arises due to its interaction with the nucleons in the Dirac sea. This is calculated using dimensional regularization scheme (see~\cite{pradip} for details). In a hot system of particles, there is a thermal distribution of real particles (on shell) which can participate in the absorption and emission process in addition to the exchange of virtual particles. The interaction of the rho with the onshell nucleons, present in the thermal bath contributes to the medium dependent part of the $\rho$-self energy. This is calculated from Eq.(\ref{pimunu}) using imaginary time formalism as, \begin{eqnarray} {\rm {Re}}\Pi_{\rm {med}}(\omega,{\bf k}\,\rightarrow\,0)& = & -\frac{16g_{VNN}^2}{\pi^2}\,\int\,\frac{p^2\,dp}{\omega_p\, (e^{\beta\,\omega_p}+1)(4\omega_p^2-\omega^2)}\nonumber\\ &&\times\,\left[\frac{1}{3}(2p^2+3M^{\ast 2})+\omega^2\left\{M^{\ast} (\frac{\kappa_V} {2M})\right.\right.\nonumber\\ &&+\left.\left.\,\frac{1}{3}(\frac{\kappa_V}{2M})^2(p^2+3M^{\ast 2})\right\} \right] \end{eqnarray} where, $\omega_p^2=p^2+M^{\ast 2}$. \section*{IV. Results} In this section we present the results of our calculation of the effective masses and decay widths of vector mesons and its effect on electromagnetic probes. For our calculations of $\rho$ and $\omega$~mesons effective masses we have used the following values of the coupling constants and masses~\cite{bonn}: $\kappa_{\rho} = 6.1,~g_{\rho NN}^2/4\pi = 0.55, m_s$= 550 MeV, $m_{\rho} = 770$ MeV, $M = 938$ MeV, $g_s^2/4\pi = 9.3$, $g_{\omega NN}^2/4\pi = 20$, and $\kappa_{\omega}$ = 0. In Fig.~3 we show the variation of the ratio of effective mass to free mass for hadrons as a function of temperature. The expected trend based on Brown - Rho scaling is also shown. Firstly, we do not observe any global scaling behaviour. Secondly, in our case the effective mass as a function of temperature falls at a slower rate. \begin{figure} \centerline{\psfig{figure=scale_new.ps,height=6cm,width=8cm}} \caption{ Ratio of effective mass to free space mass of hadrons as a function of temperature $T$.} \label{scaling} \end{figure} Many authors have investigated the issue of temperature dependence of hadronic masses within different models over the past several years. Hatsuda and collaborators~\cite{furn,adami,hatsuda} and Brown~\cite{Brown} showed that the use of QCD sum rules at finite temperature results in a temperature dependence of the $q \bar q$ condensate culminating in the following behaviour of the $\rho$-mass: \begin{equation} \frac{m_{\rho}^{\ast}}{m_{\rho}} = \left(1- \frac{T^2}{T_{\chi}^ 2}\right)^ {1/6} \end{equation} where $T_{\chi}$ is the critical temperature for chiral phase transition. Brown and Rho~\cite{rho} also showed that the requirement of chiral symmetry yields an approximate scaling relation between various effective hadron masses, \begin{equation} \frac{M^{\ast}}{M}\approx\frac{m_{\rho}^{\ast}}{m_{\rho}}\approx \frac{m_{\omega}^{\ast}}{m_{\omega}}\approx\frac{f_{\pi}^{\ast}}{f_{\pi}} \end{equation} These calculations show a dropping of hadronic mass with temperature. Calculations with non-linear $\sigma$-model, however, predict the opposite trend \cite{abijit}. \begin{figure} \centerline{\psfig{figure=rhowidth.ps,height=6cm,width=8cm}} \caption{ Rho decay width ($\Gamma_{\rho}$) as a function of temperature ($T$). Dashed and solid lines show calculations of rho width with effective mass due to nucleon loop, with and without BE. Dot dashed line represents the same but with effective mass due to pion loop. } \label{rowidth} \end{figure} The variation of the in-medium decay width ($\Gamma_{\rho}$) of the rho meson with temperature is shown in Fig.~(\ref{rowidth}). As discussed earlier, the effective mass of the rho decreases as a result of $N\bar N$ polarisation. This reduces the phase space available for the rho. Hence, we observe a rapid decrease in the rho meson width with temperature (solid line). However, the presence of pions in the medium would cause an enhancement of the decay width through induced emission. Thus when Bose Enhancement (BE) of the pions is taken into account the rho decay width is seen to fall less rapidly (dashed line); such a behaviour is observed quite clearly in ~\cite{abhee}. For the sake of completeness, we also show the variation of rho width in the case where the rho mass changes due to $\pi \pi$ loop. In this case, since the rho mass increases (though only marginally), the width increases (dot-dashed line). Now we present our results on photon emission rates from a hot hadronic gas. As discussed earlier, the variation of hadronic decay widths and masses will affect the photon spectra. The relevant reactions of photon production are $\pi \pi\,\rightarrow\,\rho\,\gamma, \pi\,\rho\,\rightarrow \pi\,\gamma, \pi\,\pi\,\rightarrow\,\eta\,\gamma$, and $\pi\,\eta\,\rightarrow\,\eta\,\gamma$ with all possible isospin combinations. Since the lifetimes of the rho and omega mesons are comparable to the strong interaction time scales, the decays $\rho\,\rightarrow\,\pi\,\pi\,\gamma$ and $\omega\,\rightarrow\,\pi^0\,\gamma$ are also included. The effect of finite resonance width of the rho meson in the photon production cross-sections has been taken into account through the propagator. \begin{figure} \centerline{\psfig{figure=newtot.ps,height=6cm,width=8cm}} \caption{ Total photon emission rate from hot hadronic matter as a function of photon energy at $T$=160, 180 MeV. The solid and dot-dashed lines show results with and without in-medium effects respectively. Inset: Total photon emission rate is plotted in linear scale as a function of photon energy in the kinematic window, $E_{\gamma}=$ 2.5 to 3.0 GeV at $T=$180 MeV. } \label{totphot} \end{figure} In Fig.~(\ref{totphot}) we display the total rate of emission of photons from hot hadronic gas including all hadronic reactions and decays of vector mesons. At $T=$180 MeV, the photon emission rate with finite temperature effects is a factor of $\sim$ 3 higher than the rate calculated without medium effects. At $T=160$ MeV the medium effects are small compared to the previous case. The transverse momentum distribution of photon has been obtained by convoluting the production rate with the space time dynamics of the system. The boost invariant hydrodynamical model of Bjorken~\cite{bjorken} has been used for the space time evolution of the system with appropriate modification due to the shift in hadronic masses~\cite{spjs}. In Fig.~\ref{wa98} we compare our results of transverse momentum distribution of photons with the preliminary results of WA98 Collaboration~\cite{wa98con}. The experimental data represents the photon spectra from Pb + Pb collisions at 158 GeV per nucleon at CERN SPS energies. The transverse momentum distribution of photons originating from the `hadronic scenario' (matter formed in the hadronic phase) with (solid line) and without(short-dash line) medium modifications of vector mesons outshine the photons from the `QGP scenario' (matter formed in the QGP phase, indicated by long-dashed line) for the entire range of $p_T$. Although photons from `hadronic scenario' with medium effects on vector mesons shine less bright than those from without medium effects for $p_T>2$ GeV, it is not possible to distinguish clearly between one or the other on the basis of the experimental data. \begin{figure} \centerline{\psfig{figure=pbpb.eps,height=6cm,width=8cm}} \caption{Total thermal photon yield in Pb + Pb central collisions at 158 GeV per nucleon at CERN SPS. The long-dash line shows the results when the system is formed in the QGP phase with initial temperature $T_i=180$ MeV at $\tau_i=1$ fm/c. The critical temperature for phase transition is taken as 160 MeV. The solid (short-dash) line indicates photon spectra when hadronic matter formed in the initial state at $T_i=230$ MeV ($T_i=270$ MeV) at $\tau_i=1$ fm/c with (without) medium effects on hadronic masses and decay widths. } \label{wa98} \end{figure} So far we have studied the effects of in-medium hadronic masses and decay widths on the photon spectra, where, the latter is found to have negligible effects. However, the modification in the rho decay width due to BE, arising from the induced emission of pions in the thermal medium, plays a crucial role in low mass lepton pair production. It has been shown in Ref.~~\cite{weldon} that this effect arises naturally in a calculation based on finite temperature field theory. For the sake of illustration, we consider the effect of thermal nucleon loop on the lepton pair production from pion annihilation via (rho mediated) vector meson dominance. In the work of Li, Ko and Brown~\cite{li}, the observed enhancement on the low mass dilepton yield was attributed solely to the the matter induced mass modification of the meson, while neglecting the BE effect. In Fig.~(\ref{ratedil}) the dilepton emission rate is plotted as a function of invariant mass at a temperature $T$=180 MeV. It is clear from the figure that the inclusion of BE effects reduces the yield by increasing the decay width. Quantitatively, a suppression by a factor of 3 is observed at the invariant mass, $M=m_{\rho}^{\ast}$. \begin{figure} \centerline{\psfig{figure=dilrate.ps,height=6cm,width=8cm}} \caption{ Dilepton emission rates as a function of invariant mass at $T$ = 180 MeV and twice nuclear matter density. Solid and dotdashed lines show the results with and without BE effects respectively. The dotted line shows the result when no medium effect is considered. } \label{ratedil} \end{figure} \section*{V. Summary and Discussions} We have discuss the general formalism to evaluate the emission of electromagnetic ejectiles from a thermal bath. This formalism has been applied to evaluate the photon and dilepton emission rate from a hot and dense hadronic system. We have calculated the effective masses and decay widths of vector mesons propagating in a hot medium. We have seen that the mass of rho meson decreases substantially due to its interactions with nucleonic excitations and it increases only marginally ($\sim$ 10-15 MeV) due to $\rho-\pi$ interactions. The overall decrease in the effective mass is due to fluctuations in the Dirac sea of nucleons with mass $M^{\ast}$ and the in-medium contribution increases very little ($\sim$ 5 -- 10 MeV) from its free space value. The omega mass drops at $T=$180 MeV by about 250 MeV from its free space value. We have evaluated the rate of photon emission from a hadronic gas of $\pi, \rho, \omega$ and $\eta$ mesons. It is seen that the photon rate increases by a factor $\sim$ 3 due to the inclusion of in-medium masses at $T=$180 MeV. We observe that the inclusion of in-medium decay widths in vector meson propagator has negligible effects on photon emission rates, although the effect of in-medium decay widths with BE is substantial for the dilepton yield. We have compared the experimental data on photon spectra from Pb + Pb collisions at energies 158 GeV per nucleon with different initial conditions. We observe that photons from hadronic scenario dominates over the photons from QGP scenario for the entire $p_T$ domain. But in the hadronic scenario the photon spectra evaluated with and without in-medium properties of vector mesons describe these data reasonably well. Hence the transverse photon spectra at present do not allow us to decide between an in-medium dropping mass and a free mass scenario. Considering the experimental uncertainty, it is not possible to state, which one, between the two is compatible with the data. Experimental data with better statistics could possibly distinguish among various scenarios. \noindent{{\bf Acknowledgement}: We thank B. Dutta-Roy, H. Gutbrod and V. Manko for useful discussions. One of us (JA) is grateful to Japan Society for Promotion of Science for financial support and Physics Department, Kyoto University where part of this manuscript was written.}\\
\section{Introduction} A measurement on a quantum mechanical system only provides partial information on the measured state. Even in the case where $N$ identical copies of the system are available, the information which can be retrieved remains bounded. This fact can be quantified using the averaged fidelity based on the following general idea. Given $N$ identical copies of a system we may consider a two-step procedure to rate the fidelity of a measuring apparatus. First, we set up a generalized quantum mechanical measurement (or positive operator valued measurement, POVM \cite{N,P}). Upon performing a measurement, its outcome provides the basis for a best guess about the incoming state. The averaged fidelity quantifies how close the final guess is from the original state averaging over the latter. For any finite number $N$ of copies of a spin $J$ pure state system the average fidelity is proven to be bounded by \cite{BM} \begin{equation} \overline {f}(N,J)={{N+1}\over {N+2 J+1}}. \label{fmax} \end{equation} The issue at stake remains to device the optimal and minimal measuring strategy for any quantum system. Explicit constructions of optimal and minimal generalized quantum mechanical measurements of spin ${1\over 2}$ systems have been presented recently in Refs. \cite{MP,H,DBE,LPT,VLPT}. The detailed construction is subtle and depends on whether the original system is in a pure or mixed state. The simplest case corresponds to measuring a spin ${1\over 2}$ system known to be in a pure state. A generalized measurement can be constructed as a resolution of the identity made with rank one hermitian operators, which are in turn built from the direct product of a given state, \begin{equation} I= \sum_{r=1}^n c^2_r | \Psi_r\rangle^N\ {}^N\langle\Psi_r|\ , \label{identity} \end{equation} where $I$ is then the identity in the maximal spin subspace. The important --and of possible future practical relevance-- result is that the maximum averaged fidelity is attained with a finite number of operators \cite{DBE}. Upon a case by case analysis, it is found that the minimum number, $n$, of such operators is a function of $N$ \begin{table}[h] \begin{center} \begin{tabular}{|c|c|c|c|c|c|} \hline $N$&1&2&3&4&5\\ \hline $n$&2&4&6&10&12\\ \hline \end{tabular} \end{center} \end{table} \noindent and is given in the table. The explicit form of Eq. (\ref{identity}) for the above cases can be found in Ref. \cite{LPT}. The far more involved case of spin ${1\over 2}$ mixed states has also been worked out in Ref. \cite{VLPT}. At variance with the pure state case, the closed expression for the maximum averaged fidelity depends on what the unbiased a priori distribution of density matrices is. Yet, explicit solutions for optimal measurements are found. Some remarkable properties emerge along the new construction. Let us briefly mention a few. Optimal measurements turn out to be structured using projectors on total spin eigenspaces and, within each eigenspace, on maximal spin component is some direction. This allows for a reuse of minimal and optimal results from the pure state case. Also, beyond two copies, some projectors are not of rank one. Explicit constructions of optimal minimal measurements are so far restricted to spin ${1\over 2}$ systems, either pure or mixed. It is the purpose of this paper to extend this analysis for arbitrary spin pure states. A number of non-trivial issues must be faced at the outset. For instance, progress in the spin ${1\over 2}$ case was triggered by the appropriate use of the Bloch vector labelling of density matrices associated to spinors. We shall resort to a similar representation in the case of arbitrary spin states, using representations of $SU(2 J+1)$. The equivalent of a Bloch vector will be shown to obey a covariant restriction. This extra work will allow for a unified general setting of the problem of optimal measurements of arbitrary spins. Finding explicit minimal optimal measurements remains a matter of case by case analysis. We shall provide explicit bounds for the minimal number of projectors, $n$, in POVMs. The case of $N=2$ will be fairly complete. Higher number of copies still need further ingenuity to get rigorous bounds. \section{ Averaged fidelity} Consider a spin $J$ particle which is in an unknown pure state $|\Psi\rangle$, \begin{equation} |\Psi\rangle = \left(\matrix{ x_1+{\rm i} y_1\cr x_2+{\rm i} y_2\cr \dots\cr x_D+{\rm i} y_D\cr }\right)\ , \label{spinor} \end{equation} where $D=2 J+1$ and the normalization of the state imposes $\sum_{i=1,..,D}\left(x_i^2+y_i^2\right)=1$. Of course, we may use a different parametrization, {\sl e.g.} \begin{equation} |\Psi\rangle = \left(\matrix{ \cos\phi\cr \sin\phi \left(x_2+{\rm i} y_2\right)\cr \dots\cr \sin\phi \left(x_D+{\rm i} y_D\right)\cr }\right)\ , \label{spinordos} \end{equation} with $0\leq \phi\leq {\pi\over 2}$ and $\sum_{i=2,..,D}\left(x_i^2+y_i^2\right)=1$. Using this second parametrization and following Ref. \cite{SAC} it is possible to prove that the volume element in the space of these states is \begin{equation} dV_D = 4\left( \sin \phi\right)^{2D-3}\ \cos\phi \ d\phi\ dS_{2D-3}\ , \label{volumeleement} \end{equation} where $dS_{2D-3}$ corresponds to the standard volume element on $S_{2 D-3}$. The total volume is \begin{equation} V_D = {4\pi^{D-1}\over (D-1)!}\ . \label{volume} \end{equation} Given $N$ identical copies of the arbitrary spin state we have \begin{equation} \vert\Psi\rangle^N\equiv \vert\Psi\rangle \otimes\vert\Psi\rangle \otimes...{}^N...\otimes\vert\Psi\rangle. \label{psiN} \end{equation} A measurement on this enlarged system will bring richer information on $\vert\Psi\rangle$ than $N$ separate measures on its respective copies \cite{PW}. Setting a generalized quantum measurements consists in providing a resolution of the identity of the type \begin{equation} \sum^n_{i=1}\ c^2_r\ \vert\Psi_r \rangle^{N}{}^{N}\langle\Psi_r\vert+ P_N= I\ , \label{povm} \end{equation} where $P_N$ is the projector on the space different from the one spanned from states of the form given in Eq. (\ref{psiN}). We already have all the necessary elements to define and compute the averaged fidelity. Upon measuring $\vert\Psi\rangle^N$ with the above POVM, a given outcome labelled by $r$ will result with probability $\vert ^N\langle\Psi\vert\Psi_r\rangle^N\vert^2$. The natural guess for the initial pure state is, then, $\vert \Psi_r\rangle$ (this is only the best strategy if the initial state is known to be pure; the best guess for a mixed state is not the same state as the outcome of the POVM \cite{VLPT}). The overlap of this guess with the original state is just $\vert \langle\Psi\vert\Psi_r\rangle\vert^2$. The averaged or mean fidelity is defined as the product of the probability for $r$ being triggered times the overlap between the ensuing guess and the original state, averaged over all possible initial unknown states, \begin{equation} \overline f(N,J)\equiv {1\over V_{2 J+1}} \sum_{r=1}^n c_r^2 \int_0^{\pi\over 2} d\phi\ \left(\sin\phi\right)^{4 J-1}\cos\phi \int dS_{4J-1} \left| ^N\langle\Psi\vert\Psi_r\rangle^N\right|^2 \ \left| \langle\Psi\vert\Psi_r\rangle\right|^2\ . \label{averagefid} \end{equation} To evaluate the above expression it is convenient to use the freedom to choose the integration variables to set each individual $\vert \Psi_r\rangle$ as a spinor with only nonvanishing first component. Then, \begin{equation} \overline f(N,J)= {1\over V_{2 J+1}} \sum_{r=1}^n c_r^2 \int_0^{\pi\over 2} d\phi\ \left(\sin\phi\right)^{4 J-1} \left(\cos\phi\right)^{2 N+3} S_{4 J-1}\ . \label{paso} \end{equation} We finally get \begin{equation} \overline f(N,J)= {(2 J)! (N+1)!\over (2 J+N+1)!}\ \sum_{r=1}^n c_r^2\ . \label{pasodos} \end{equation} This sum is easily calculated. It is just the dimension of the space spanned by the totally symmetric tensor of order $N$ whose indices can take $2J+1$ values, \begin{equation} \sum_{r=1}^n c_r^2 = {(2 J+N)!\over N! (2J)!}\ . \label{sumacr} \end{equation} Thus, \begin{equation} \overline f(N,J)= {N+1\over N+2J+1}\ , \label{averagefidagain} \end{equation} which corresponds to Eq. (\ref{fmax}) and was obtained in Ref. \cite{BM} using different techniques. \section{Generalized Bloch form of arbitrary spin pure states} It is sometimes uselful to represent the state of a spin ${1\over 2}$ system using the Bloch representation, \begin{equation} \rho = {1\over 2} I+ {1\over 2} \vec b \cdot \vec \sigma \ , \label{blochform} \end{equation} where $\vec b$ is a vector living within the unit sphere. Pure states correspond to the surface of the sphere, that is $\vec b^2=1$. A similar but more complicated construction is possible for arbitrary spin particles. Consider a pure state of a spin $J$ particle. One may represent it using {\sl e.g.} Eq. (\ref{spinor}). Alternatively we may construct its associated density matrix and write \begin{equation} \rho= {1\over 2 J+1}I + \sqrt{J\over 2 J+1}\ n_a \lambda_a \qquad a=1,\dots,4J(J+1) \ , \label{genbloch} \end{equation} where $\lambda_a$ are the generators of the $SU(2J+1)$ normalized by \begin{equation} {\rm Tr}\left(\lambda_a \lambda_b\right) = 2 \delta_{ab}\ , \label{normlambdas} \end{equation} and $\hat n$ is the normalized vector that plays the role of a generalized Bloch vector. The coefficients in Eq. (\ref{genbloch}) are chosen in such a way that $ {\rm Tr} \rho={\rm Tr} \rho^2 = 1$. A simple counting of degrees of freedom shows that a spin $J$ pure state is described by $4J$ real parameters whereas the generalized Bloch vector carries $4J(J+1)-1$. A mismatch appears for $J>{1\over 2}$ which implies that severe constraints must limit the subspace of valid vectors $\hat n$. Indeed, pure states must verify $\rho=\rho^2$ which translates into \begin{equation} d_{abc} n_a n_b = {2J-1\over\sqrt{ J(2J+1)}} n_c \ , \label{compact} \end{equation} when Eq. (\ref{genbloch}) is used and where $d_{abc}$ are the completely symmetric symbols associated to $SU(2 J+1)$, defined through the anticommutator fo the generators of the group \cite{PT} \begin{equation} \{ \lambda_a,\lambda_b\} = {4\over 2J+1} \delta_{ab} I + 2 \ d_{abc} \lambda_c \label{defd} \end{equation} which verify \begin{equation} d_{abb}=0\quad,\quad d_{abc} d_{dbc} = {(2J-1)(2J+3)\over 2J+1} \delta_{ad}\ . \label{dprop} \end{equation} Some useful properties of the vectors $\hat n$ follow from the above general covariant constraint (\ref{compact}), \begin{eqnarray} && d_{abc} n_a n_b n_c = {2 J-1\over \sqrt{J(2J+1)}}\ , \nonumber\\ && d_{abe} d_{cde} n_a n_b n_c n_d = { (2 J-1)^2\over J(2J+1)} \ , \label{dnnn} \end{eqnarray} where it is clear that for spin $J={1\over 2}$ the simple structure of $SU(2)$ makes the $d$-symbols to vanish and the r.h.s. to be identically zero. We can also deduce the useful constraint which follows from the positivity of the square of the scalar product of two arbitrary spin $J$ pure states which reads \begin{equation} \vert\langle \Psi\vert\Psi'\rangle\vert^2 = {\rm Tr}\left(\rho \rho'\right)= {1\over 2J+1} \left(1+ 2 J \hat n\cdot\hat n'\right) \geq 0\ . \label{scalarproduct} \end{equation} Generalized Bloch vectors are thus constrained to have scalars products bounded by \begin{equation} \hat n\cdot \hat n' \geq -{1\over 2J}\ . \label{boundscalarproduct} \end{equation} Two pure states are orthogonal then when the scalar product of their generalized Bloch vectors satisfies the equality in Eq. (\ref{boundscalarproduct}). Let us illustrate the construction of a Bloch vector for the $J=1$ example. In this case, the density matrix representing the system can be connected to the standard spinor-like representation. For instance, taking $J=1$ it is easy to see that the generalized Bloch vector corresponds to Eq. (\ref{spinor}) if \begin{eqnarray} n_1&=&\sqrt{3} \left(x_1 x_2+y_1y_2\right) \qquad n_2=\sqrt{3} \left(x_1 y_2-x_2y_1\right)\nonumber \\ n_4&=&\sqrt{3} \left(x_1 x_3+y_1y_3\right) \qquad n_5=\sqrt{3} \left(x_1 y_3-x_3y_1\right)\nonumber \\ n_6&=&\sqrt{3} \left(x_2 x_3+y_2y_3\right) \qquad n_7=\sqrt{3} \left(x_2 y_3-x_3y_2\right)\nonumber \\ n_3&=&{\sqrt{3}\over 2} \left(x_1^2+y_1^2-(x_2^2+y_2^2)\right) \qquad n_8={1\over 2} \left(1- 3(x_3^2+y_3^2)\right) \label{explicitbloch} \end{eqnarray} and $\lambda_a$ are taken in the Gell-Mann representation of $SU(3)$ \cite{PT}. Note that symmetric and antisymmetric combinations of the spinor components build the raising and lowering generators, whereas the Casimir combinations correspond to diagonal ones. Generalization of this construction for arbitrary spin $J$ based on the $SU(2J+1)$ group is straigthforward. The advantage of using a generalized Bloch representation for arbitrary spin pure states will become apparent shortly, when all our equations will be manifestly $SU(2J+1)$ covariant and real. This is equivalent to note that the difference between working with spinors, which live in the fundamental representation of the group, or with Bloch vectors, which live in the adjoint representation, is that the second is real. \section{Optimal measurements for a single copy of a system} Let us go back to the construction of a generalized quantum measurement of arbitrary spin systems. We basically need to solve for the minimal set of $\vert\Psi_r\rangle$ states such that Eq. (\ref{povm}) is fulfilled. We have found convenient to project out the $P_N$ piece using \begin{equation} \sum^n_{r=1} \ c^2_r\ \left| ^N\langle\Psi\vert\Psi_r\rangle^N\right|^2 =1\qquad \forall \vert\Psi\rangle\ \ , \label{basiceq} \end{equation} This equation can be also written in the Bloch representation as \begin{equation} \sum^n_{r=1} \ c^2_r {1\over (2J+1)^N} \left(1 + 2J\sum_{a}\ n_a n_a(r) \right)^N = 1\ , \label{blocheq} \end{equation} where every $\hat n(r)$ corresponds to a pure state in the POVM and $\hat n$ to the original pure state. It is clear that the simplest situation we may face corresponds to having a single copy of the unknown state. The optimal and minimal measurement for such a case is, of course, known to correspond to a von Neumann measurement. We shall, though, proceed in a more general way and set the {\sl modus operandi} for the more elaborate cases as deviced in Ref. \cite{LPT}. The equation (\ref{basiceq}) with $N = 1$ can be demonstrated (with a little effort) to be equivalent to \begin{eqnarray} &&\sum^n_{r=1} \ c^2_r\ \left( x_j(r) x_k(r)+y_j(r)y_k(r)\right)=\delta_{jk}\nonumber\\ &&\sum^n_{r=1} \ c^2_r\ \left( x_j(r) y_k(r)-x_k(r)y_j(r)\right)=0 \qquad j,k=1,\dots,2J+1 \label{previouseq} \end{eqnarray} Using the insight given by Eq. (\ref{blocheq}) and the result of Eq. (\ref{sumacr}), this set of $(2J+1)^2$ independent equations can be rewritten in terms of the Bloch vector as \begin{eqnarray} && \sum^n_{r=1} \ c^2_r\ = 2J+1 \nonumber\\ && \sum^n_{r=1} \ c^2_r\ n_a(r) =0\ , \label{firstset} \end{eqnarray} where it is important to remember the constraints limiting $\hat n(r)$. For instance, scalar products between any pair $\hat n(r)\cdot\hat n(s) \geq -{1\over 2J}$, thus \begin{equation} \sum_{r\not= s} c_r^2 \left({1\over 2J}+ \hat n(r)\cdot\hat n(s) \right) \geq 0\ . \label{basicnone} \end{equation} Using the set of equations (\ref{firstset}) the above inequality can be transformed into \begin{equation} 1- c_s^2\geq 0\qquad \forall s=1,\dots,n \label{cscosntraint} \end{equation} Suming over all $s$ we get \begin{equation} n\geq 2 J+1 \label{none} \end{equation} This bound is indeed saturated by a von Neumann measurement, that is \begin{eqnarray} n_{min}&=&2 J+1\nonumber\\ c_s^2=1\ \ \forall s\quad \quad &,&\quad \hat n(r)\cdot\hat n(s)=-{1\over 2J}\ \ \forall r\not= s \label{nonesol} \end{eqnarray} The explicit standard construction for $J=1$ is recovered as the solution to this $N=1$ POVM \begin{equation} \vert\Psi_1\rangle= \left(\matrix{1\cr 0\cr 0\cr}\right)\quad , \quad \vert\Psi_2\rangle= \left(\matrix{0\cr 1\cr 0\cr}\right)\quad , \quad \vert\Psi_3\rangle= \left(\matrix{0\cr 0\cr 1\cr}\right)\ . \label{explicitnoneone} \end{equation} Or, alternatively, \begin{eqnarray} &&\hat n(1)=\left(0,0,{\sqrt{3}\over 2},0,0,0,0,{1\over 2}\right)\nonumber\\ &&\hat n(2)=\left(0,0,-{\sqrt{3}\over 2},0,0,0,0,{1\over 2}\right)\nonumber\\ &&\hat n(3)=\left(0,0,0,0,0,0,0,-1\right) \label{explicitonetwo} \end{eqnarray} We are now in a position to appreciate the advantage of resorting to a Bloch-like parametrization. It is easier to deal with Eq. (\ref{firstset}) than with Eq. (\ref{previouseq}). The use of $\hat n(r)$ introduces a simple covariant, yet constrained, formulation. Some extra subtleties will play a relevant role in the more complicated cases. \section{Optimal measurements for the $N=2$ case} Let us face the case where $N=2$ identical copies of the system are at our disposal. Following the same reasoning as before we start by writing Eq. (\ref{basiceq}) in terms of the basic spinor representation. This leads to \begin{eqnarray} &&\sum_{r=1}^n c_r^2 \ \left(x_i(r)x_j(r)+y_i(r) y_j(r)\right) \left(x_k(r)x_l(r)+y_k(r) y_l(r)\right)={1\over 4}\left( 2 \delta_{ij}\delta_{kl}+\delta_{ik}\delta_{jl}+\delta_{il}\delta_{jk} \right)\nonumber\\ &&\sum_{r=1}^n c_r^2 \ \left(x_i(r)y_j(r)-x_j(r) y_i(r)\right) \left(x_k(r)y_l(r)-x_l(r) y_k(r)\right)={1\over 4}\left(\delta_{ik} \delta_{jl}-\delta_{il}\delta_{jk}\right) \nonumber\\ &&\sum_{r=1}^n c_r^2 \ \left(x_i(r)x_j(r)+y_i(r) y_j(r)\right) \left(x_k(r)y_l(r)-x_l(r) y_k(r)\right)=0 \label{secondprevious} \end{eqnarray} The system is now quadratic in the basic structures appearing linearly in the $N=1$ case. Using the Bloch vector representation, these $(2 J+1)^2 (2J^2+2J+1)$ equations can be recast into \begin{eqnarray} &&\sum^n_{r=1} \ c^2_r \ = (2J+1)(J+1)\equiv B \nonumber\\ &&\sum^n_{r=1} \ c^2_r \ n_a(r)=0 \nonumber\\ &&\sum^n_{r=1} \ c^2_r \ n_a(r) n_b(r) =B\ {1\over 4J(J+1)}\ \delta_{ab} \label{secondset} \end{eqnarray} A general pattern is emerging. Higher $N$ optimal measurements demand a finer grained resolution of the identity. The Bloch vectors are required to satisfy isotropy conditions in $SU(2 J+1)$ group space. The determination of the factor ${1 \over 4J(J+1)}$ has been done using that $\hat n$ is a normalized vector and Eq. (\ref{sumacr}). It is easy to verify that the set of equations (\ref{secondset}) provides a solution for Eq. (\ref{blocheq}). From the above basic set of equations it is easy to get \begin{eqnarray} &&\sum^n_{r\not= s} \ c^2_r \ = B -c_s^2 \nonumber\\ &&\sum^n_{r\not= s} \ c^2_r \ \hat n(r)\cdot \hat n(s)=-c_s^2 \nonumber\\ &&\sum^n_{r\not= s} \ c^2_r \ (\hat n(r)\cdot \hat n(s))^2 =B\ {1\over 4J(J+1)}- c_s^2\ \label{secondsetbis} \end{eqnarray} Then we may argue that \begin{equation} \sum_{r\not= s} \ c_r^2\ \left(b + \hat n(r)\cdot \hat n(s)\right)^2 \geq 0\ . \label{trick} \end{equation} which is extremized by $b={c_s^2\over B-c_s^2}$ leading to \begin{equation} n\geq (2J+1)^2 \qquad,\qquad c_s^2\leq {J+1\over 2J+1} \ \ \forall s \label{boundsntwo} \end{equation} For $J={1\over 2}$ this bound agrees with the known solution of the tetrahedron (see Introduction and Ref. \cite{LPT}), and generalizes it in the following sense. The solution $n=(2 J+1)^2$ also forces all scalar products to be $\hat n(r)\cdot\hat n(s)=-{1\over 4 J(J+1)}$. This corresponds to a hypertetrahedron in $(2J+1)^2-1$ dimensions, exactly those of the adjoint representation of $SU(2 J+1)$. Let us just write the explicit solution for $J=1$ \begin{eqnarray} &&\hat n(1)=\left({1\over 2},{\sqrt{3}\over 2},0,0,0,0,0,0\right)\nonumber\\ &&\hat n(2)=\left({1\over 2},-{\sqrt{3}\over 4},{3\over 4},0,0,0,0,0\right)\nonumber\\ &&\hat n(3)=\left({1\over 2},-{\sqrt{3}\over 4},-{3\over 4},0,0,0,0,0\right)\nonumber\\ &&\hat n(4)=\left(-{1\over 4},0,0,0,{\sqrt{6}\over 4},{\sqrt{3}\over 4},-{\sqrt{6}\over 4},0\right)\nonumber\\ &&\hat n(5)=\left(-{1\over 4},0,0,{3\sqrt{2}\over 8},-{\sqrt{6}\over 8} ,{\sqrt{3}\over 4},{\sqrt{6}\over 8},-{3\sqrt{2}\over 8}\right)\nonumber\\ &&\hat n(6)=\left(-{1\over 4},0,0,-{3\sqrt{2}\over 8},-{\sqrt{6}\over 8} ,{\sqrt{3}\over 4},{\sqrt{6}\over 8},{3\sqrt{2}\over 8}\right)\nonumber\\ &&\hat n(7)=\left(-{1\over 4},0,0,0,{\sqrt{6}\over 4},-{\sqrt{3}\over 4},{\sqrt{6}\over 4},0\right)\nonumber\\ &&\hat n(8)=\left(-{1\over 4},0,0,-{3\sqrt{2}\over 8},-{\sqrt{6}\over 8} ,-{\sqrt{3}\over 4},-{\sqrt{6}\over 8},-{3\sqrt{2}\over 8}\right)\nonumber\\ &&\hat n(9)=\left(-{1\over 4},0,0,{3\sqrt{2}\over 8},-{\sqrt{6}\over 8} ,-{\sqrt{3}\over 4},-{\sqrt{6}\over 8},{3\sqrt{2}\over 8}\right) \label{explicittwotwo} \end{eqnarray} There is still the need to perform the non-obvious step of finding out whether this solution does correspond to a set of spin 1 states. For completeness we give this final form of the solution, that is the explicit states $\vert\Psi_1\rangle$ through $\vert\Psi_9\rangle$ which form the POVM, \begin{equation} \vert\Psi_1\rangle= \left(\matrix{1\cr 0\cr 0\cr}\right)\quad \vert\Psi_2\rangle= \left(\matrix{{1\over 2}\cr {\sqrt{3}\over 2}\cr 0\cr}\right)\quad \vert\Psi_3\rangle= \left(\matrix{{1\over 2}\cr -{\sqrt{3}\over 2}\cr 0\cr}\right) \nonumber \label{explicitketsone} \end{equation} \begin{equation} \vert\Psi_4\rangle= \left(\matrix{{1\over 2}\cr {\rm i} {1\over 2}\cr {1\over \sqrt{2}}\cr}\right)\quad \vert\Psi_5\rangle= \left(\matrix{{1\over 2}\cr {\rm i}{1\over 2}\cr -{1\over 2 \sqrt{2}} +{\rm i}{\sqrt{3}\over 2\sqrt{2}}\cr}\right)\quad \vert\Psi_6\rangle= \left(\matrix{{1\over 2}\cr {\rm i}{1\over 2}\cr -{1\over 2 \sqrt{2}}- {\rm i}{\sqrt{3}\over 2\sqrt{2}}\cr}\right) \nonumber \label{explicitketstwo} \end{equation} \begin{equation} \vert\Psi_7\rangle= \left(\matrix{{1\over 2}\cr - {\rm i}{1\over 2}\cr {1\over \sqrt{2}}\cr}\right)\quad \vert\Psi_8\rangle= \left(\matrix{{1\over 2}\cr - {\rm i}{1\over 2}\cr -{1\over 2 \sqrt{2}}+{\rm i} {\sqrt{3}\over 2\sqrt{2}}\cr}\right)\quad \vert\Psi_9\rangle= \left(\matrix{{1\over 2}\cr - {\rm i}{1\over 2}\cr -{1\over 2 \sqrt{2}}-{\rm i} {\sqrt{3}\over 2\sqrt{2}}\cr}\right) \label{explicitketsthree} \end{equation} Note that all the spinors have scalar products with modulus equal to ${1\over 2}$. \section{Optimal measurements for the $N=3$ case} The systematics of our approach are already set. It is, though, in the case of three copies where a major difference between spin ${1\over 2}$ and higher spin systems appears. Following an analogous reasoning to the one in the previous sections we get \begin{eqnarray} &&\sum^n_{r=1} \ c^2_r \ = {(2J+3)!\over 3! (2J)!}\equiv C \nonumber\\ &&\sum^n_{r=1} \ c^2_r \ n_a(r)=0 \nonumber\\ &&\sum^n_{r=1} \ c^2_r \ n_a(r) n_b(r) =C\ {1\over 4J(J+1)}\ \delta_{ab}\nonumber\\ &&\sum^n_{r=1} \ c^2_r \ n_a(r) n_b(r) n_c(r) =C\ {1\over 4J(J+1) (2 J+3)}\ \left({2J+1\over J}\right)^{1\over 2} d_{abc} \label{thirdset} \end{eqnarray} We have used Eqs. (\ref{sumacr}), (\ref{dprop}) and (\ref{dnnn}) for determining the factor ${1\over 4J(J+1)(2 J+3)}\left({2J+1\over J}\right)^{1\over 2}$. Again it is easy to prove that Eq. (\ref{thirdset}) verify Eq. (\ref{blocheq}). For the first time the r.h.s. of one of the equations displays a tensor structure based on the $d$-symbol. Such a term would vanish for $J={1\over 2}$ due to the simpler structure of $SU(2)$, but is expected for higher spins (note that the conditions (\ref{dnnn}) are zero for spin $1\over 2$). A bound on the number of projectors appearing in a optimal POVM can be obtained following the by now standard procedure of investigating manifestly positive combinations. In this case, starting from \begin{equation} \sum_{r\not= s} \left({1\over 2J}+\hat n(r)\cdot \hat n(s)\right) \left( b+\hat n(r)\cdot \hat n(s)\right)^2 \geq 0\ , \label{startnthree} \end{equation} one gets \begin{equation} n\geq (J+1) (2 J+1)^2 \label{nthreebound} \end{equation} and $c_s^2\leq {(2J+3)\over 3(2J+1)}$. That is, $n\geq 6$ for spin ${1\over 2}$ (which agrees with the known result in Ref. \cite{LPT}), $n\geq 18$ for spin 1, $n\geq 40$ for spin ${3\over 2}$, etc. Saturating this bound is impossible for certain cases as implied by the following simple argument. If the bound were to be saturated, then Eq. (\ref{startnthree}) becomes a restricting condition for all scalar products. Indeed, $\hat n(r)\cdot \hat n(s)$ is either $-{1\over 2J}$ or else ${2J-1\over 2 J (2J+3)}$ for any pair $r\not= s$. If we fix any $s$ and assume that the minimal solution carries $p$ scalar products of the first type and q of the second, it follows that Eq. (\ref{thirdset}) imposes $p={1\over 2}J(2J+1)^2$ and $q={1\over 2} J(2J+3)^2$. For any $J$ half integer or even this causes no problem but for odd integer values of the spin this leads to non-integer pairs, which is absurd. Thus, in such a case, the bound cannot be saturated. \section{Conclusions} We have presented explicit solutions for minimal optimal POVMs acting on arbitrary spin $J$ systems for the case when two copies are available. For $N=3$ we have provided a rigorous bound. The key idea to simplify the analysis consists in using Bloch representation for pure arbitrary spin states. These vectors do not span a naive $(2J+1)^2-1$ sphere, but rather an intrincate subspace defined through covariant restrictions. The power of such covariance makes the set of equations simple \begin{eqnarray} &&\sum^n_{r=1} \ c^2_r \ = {(2J+N)!\over N! (2J)!} \nonumber\\ &&\sum^n_{r=1} \ c^2_r \ n_a(r)=0 \nonumber\\ &&\sum^n_{r=1} \ c^2_r \ n_a(r) n_b(r) = {(2J+N)!\over N! (2J)!}\ {1\over 4J(J+1)}\ \delta_{ab}\nonumber\\ &&\sum^n_{r=1} \ c^2_r \ n_a(r) n_b(r) n_c(r) = {(2J+N)!\over N! (2J)!}\ {1\over 4J(J+1) (2 J+3)}\ \left({2J+1\over J}\right)^{1\over 2} d_{abc} \nonumber\\ &&\dots \label{nset} \end{eqnarray} In order to analyzed a given case with $N$ copies of the spin $J$ particle, it is necessary to retain \begin{equation} {(4J(J+1)+N)!\over N! (4J(J+1)!} \label{numbereq} \end{equation} equations in the system, that is as many rows in Eq. (\ref{nset}) as $N+1$. Our results confirm the expected increase of needed projectors to build a POVM as the spin of the system increases. The instances analyzed, that is $N=1,2,3$, seem to point at a dependence of the type \begin{equation} n_{min} \sim J^N\ . \label{generaldep} \end{equation} \vspace*{1cm} \section{Acknowledgments} We are grateful to R. Tarrach and G. Vidal for continously sharing their insight with us. Financial support from CICYT, contract AEN95-0590, and from CIRIT, contract 1996GR00066 are acknowledged. A. A. acknowledges a grant from MEC.
\section{Introduction} The solutions of Yang-Baxter equation (YBE) depending on spectral parameter are of special importance in mathematical physics. They are connected with Wess-Zumino-Witten models and affine Toda field theories \cite{BERN},\cite{BERN1}. The algebraic base of these theories is formed by the universal enveloping algebra $U(g[u])$ of polynomial current algebra $g[u]$ with $g$ being a finite-dimensional complex Lie algebra. Yangians $Y(g)$ were introduced by Drinfeld \cite{DRIN1} as a quantum deformations of the algebra $U(g[u])$. They correspond to the rational solutions of the classical Yang-Baxter equation (CYBE) first found by Belavin \cite{BEL} and completely described by Belavin and Drinfeld \cite{BELDR}. The classification of rational solutions of CYBE for simple Lie algebras was performed by Stolin \cite{STO}. The quantum deformation ${\cal R}(\lambda)$ of such rational solution \begin{equation} \label{rat-sol} r(\mu - \nu) = \frac{C^2}{\mu - \nu} \end{equation} (with $C^2$ being the second order Casimir element of the algebra g) satisfies the parametric Yang-Baxter equation, $$ {\cal R}_{12}(\lambda_1 - \lambda_2){\cal R}_{13}(\lambda_1 - \lambda_3) {\cal R}_{23}(\lambda_2 - \lambda_3) = {\cal R}_{23}(\lambda_2 - \lambda_3){\cal R}_{13}(\lambda_1 - \lambda_3) {\cal R}_{12}(\lambda_1 - \lambda_2), $$ and realizes the morphism to the Yangian with the opposite comultiplication: \begin{equation} \label{pqybe} \left( T_{\lambda} \otimes {\rm id} \right) \Delta^{\rm op}(a) = {\cal R}(\lambda) \left( (T_{\lambda} \otimes {\rm id}) \Delta(a) \right) \left( {\cal R}(\lambda) \right)^{-1}, \end{equation} where $T_{\lambda}$ is a parameter shifting operator. In this report we consider the problem of constructing the quantizations of $U(a[u])$ where the algebra $a$ is not semisimple and also present rational solutions of CYBE for such algebras. Our approach is based on studying the properties of the boundaries of parametrized sets of Yangians. We find that under certain conditions the corresponding algebraic constructions survive on the boundaries of the parametrized domain and can be explicitely described. To reach the boundaries we use the restricted limiting procedure based on the existence of dual structure in the corresponding Hopf algebra $H$. The existence of a dual structure in $H$ means that there exists the two-dimensional set of Hopf algebras containing $H$. One of the properties of such sets is that when the corresponding parameters go to its limiting values the algebraic construction survives. In most of the cases the Hopf algebra thus obtained is inequivalent to the original one. This was demonstrated for deformation quantizations of the finite-dimensional Lie algebras \cite{LYATKA}. Now we show that the same is true also for Yangians and other possible quantizations of $U(g[u])$. \section{Yangian $Y(sl(2))$ and its natural limits} The Yangian $Y(sl(2))$ is a Hopf algebra generated by the elements $ \{ e_k, h_k, f_k \} $ $ (k \in {\bf Z}_+ ) $ with relations \begin{equation} \label{ini-my} \begin{array}{l} \begin{array}{lcl} \, [h_k , h_l ] = 0, & & [e_k , f_l ] = h_{k+l},\\[1mm] \, [h_0 , e_l ] = 2e_{l}, & & [h_0 , f_l ] = -2f_{l},\\[1mm] \end{array} \\ \begin{array}{lcl} \, [h_{k+1} , e_l ] -[h_k , e_{l+1} ] & = & \hbar \{ h_k,e_l \},\\[1mm] \, [h_{k+1} , f_l ] -[h_k , f_{l+1} ] & = & -\hbar \{ h_k,f_l \},\\[1mm] \, [e_{k+1} , e_l ] -[e_k , e_{l+1} ] & = & \hbar \{ e_k,e_l \},\\[1mm] \, [f_{k+1} , f_l ] -[f_k , f_{l+1} ] & = & -\hbar \{ f_k,f_l \},\\ \end{array} \end{array} \end{equation} where $\hbar$ is the deformation parameter. The coproducts for the generators of $sl(2) \in Y(sl(2))$ rest primitive: \begin{equation} \label{ini-py} \Delta (x) = x \otimes 1 + 1 \otimes x; \, \, \, x \in sl(2) , \end{equation} while the nontrivial coalgebraic part is uniquely defined by the following comultiplications (and the multiplications (\ref{ini-my}) above): \begin{equation} \label{ini-cy} \begin{array}{lcl} \Delta (e_1) & = & e_1 \otimes 1 + 1 \otimes e_1 + \hbar h_0 \otimes e_0, \\ \Delta (f_1) & = & f_1 \otimes 1 + 1 \otimes f_1 + \hbar f_0 \otimes h_0. \\ \end{array} \end{equation} To be able to study general properties of Yangian's Hopf structure the generating functions formalism is especially convenient. In terms of \begin{equation} \label{gen-fs} \begin{array}{l} e(u) := \sum_{k \geq 0} e_k u^{-k-1}, \;\; f(u) := \sum_{k \geq 0} f_k u^{-k-1}, \\[1mm] h(u) :=1 + \hbar\chi(u) := 1 + \hbar \sum_{k \geq 0} h_k u^{-k-1} \end{array} \end{equation} the compositions generated by (\ref{ini-my},\ref{ini-py},\ref{ini-cy}) look like \begin{equation} \label{ini-myg} \begin{array}{lcl} \,[h(u) , h(v)] = 0, & & [e(u), f(v)] = - \frac{1}{\hbar}\frac{h(u) - h(v)}{u-v} ,\\[2mm] \,[h(u), e(v)] = - \hbar\frac{ \{ h(u),(e(u) - e(v))\} }{u-v} , & & [h(u), f(v)] = \hbar\frac{ \{ h(u),(f(u) - f(v))\} }{u-v} ,\\[2mm] \,[e(u), e(v)] = - \hbar\frac{ (e(u) - e(v))^2 }{u-v} , & & [f(u), f(v)] = \hbar\frac{ (f(u) - f(v))^2 }{u-v} . \end{array} \end{equation} The coproducts for the generating functions are written in the form proposed by Molev \cite{MOL} \begin{equation} \label{ini-cyg} \begin{array}{lcl} \Delta (e(u)) & = & e(u) \otimes 1 + \sum^{\infty}_{k=0} (-1)^k \hbar^{2k} (f(u + \hbar))^k h(u) \otimes (e(u))^{k+1},\\[3mm] \Delta (f(u)) & = & 1 \otimes f(u) + \sum^{\infty}_{k=0} (-1)^k \hbar^{2k} (f(u))^{k+1} \otimes h(u)(e(u + \hbar))^k,\\[3mm] \Delta (h(u)) & = & \sum^{\infty}_{k=0} (-1)^k (k+1) \hbar^{2k} (f(u + \hbar))^{k} h(u) \otimes h(u)(e(u + \hbar))^{k}. \end{array} \end{equation} The dual structure in $Y(sl(2))$ is inherited from the basic subalgebra $sl(2)$ interpreted as a classical double. When the double structure is in its prime form the dual parameters and the corresponding analytical family of Hopf algebras can be canonically introduced (see \cite{LYM}). In the case of Yangian this double is factorized. To get the necessary parametrization we must reconstruct the prefactorized relations. Such reconstruction can be achieved if one takes into account that the unfactorized classical double of the initial $sl(2)$ subalgebra has the form \begin{equation} \label{c-dub} \begin{array}{lcl} \,[h, h'] = 0, & & [e, f] = \frac{1}{2} (h + h'), \\ \,[h, e] = 2 e, & & [h, f] = -2 f, \\ \,[h', e] = 2 e, & & [h', f] = -2 f. \end{array} \end{equation} This means that the transformation \begin{equation} \label{subs} \begin{array}{lcl} e_k & \rightarrow & \frac{1}{p} e_k, \\[1mm] f_k & \rightarrow & \frac{1}{t} f_k, \\[1mm] h_k & \rightarrow & \frac{1}{2}(\frac{h_k}{p} + \frac{h'_k}{t}) \end{array} \end{equation} accompanied by the rescaling of the original $sl(2)$ structure constants will lead to the parametrization that would not be canonical in the whole two-dimensional domain but might work well just in the neighborhood of its boundaries. According to the general rules \cite{LYM} the deformation parameter must be also rescaled: \begin{equation} \hbar \rightarrow pt. \end{equation} All these transformations and rescalings produce the pa\-ra\-met\-rized algebraic construction $Y_{pt}(sl(2))$ well defined only when one of the parameters ($p$ or $t$) is small. When a deformation quantization algebra admits a canonical parametrization (according to some dual structure) its quasiclassical limits can be found. In our case the limits $ Y_{p,0}(a) := \lim_{t \rightarrow 0}Y_{pt}(sl(2))$ and $Y_{0,t}(b) := \lim_{p \rightarrow 0}Y_{pt}(sl(2))$ are equivalent due to the selfduality of the Borel subalgebra in the classical double algebra (\ref{c-dub}). Thus we need to consider only one of these algebraic constructions, that we call boundary Yangians. The limiting procedure gives the following structure constants for the boundary Yangian $Y_{p,0}(a)$: \begin{equation} \label{lim-my} \begin{array}{l} \begin{array}{lll} \, [h_k , h_l ] = 0, & [h'_k , h'_l ] = 0, & [h_k , h'_l ] = 0,\\[1mm] \, [h_0 , e_l ] = 4pe_{l}, & [h_0 , f_l ] = -4pf_{l}, & \\[1mm] \, [h'_0 , e_l ] = 0, & [h'_0 , f_l ] = 0, & [e_k , f_l ] = \frac{p}{2} h'_{k+l}, \\[1mm] \end{array}\\[2mm] \begin{array}{lll} \, [h_{k+1} , e_l ] -[h_k , e_{l+1} ] & = & p^2 \{ h'_k,e_l \},\\[1mm] \, [h_{k+1} , f_l ] -[h_k , f_{l+1} ] & = & -p^2 \{ h'_k,f_l \},\\[1mm] \, [h'_{k+1} , e_l ] -[h'_k , e_{l+1} ] & = & 0,\\[1mm] \, [h'_{k+1} , f_l ] -[h'_k , f_{l+1} ] & = & 0,\\[1mm] \, [e_{k+1} , e_l ] -[e_k , e_{l+1} ] & = & 0,\\[1mm] \, [f_{k+1} , f_l ] -[f_k , f_{l+1} ] & = & 0,\\[1mm] \end{array} \end{array} \end{equation} The coproducts for the "zero mode" generators $e_0,f_0,h_0,h'_0$ rest primitive. The others are defined by the relations: \begin{equation} \label{lim-cy} \begin{array}{lcl} \Delta (e_1) & = & e_1 \otimes 1 + 1 \otimes e_1 + \frac{p}{2} h'_0 \otimes e_0, \\[1mm] \Delta (f_1) & = & f_1 \otimes 1 + 1 \otimes f_1 + \frac{p}{2} f_0 \otimes h'_0, \\[1mm] \Delta (h_1) & = & h_1 \otimes 1 + 1 \otimes h_1 + \frac{p}{2} (h'_0 \otimes h_0 + h_0 \otimes h'_0) \\[1mm] & & - 4p f_0 \otimes e_0,\\[1mm] \Delta (h'_1) & = & h'_1 \otimes 1 + 1 \otimes h'_1 + \frac{p}{2}(h'_0 \otimes h'_0). \\[1mm] \end{array} \end{equation} The internal structure of this Hopf algebra becomes more transparent in terms of generating functions: \begin{equation} \label{lim-myg} \begin{array}{l} \begin{array}{lll} \, [\chi(u) , \chi(v) ] = 0, & [\chi'(u) , \chi'(v) ] = 0, & [\chi(u) , \chi'(v) ] = 0,\\[1mm] \, [\chi'(u) , e(v) ] = 0, & [\chi'(u) , f(v) ] = 0, & \\[2mm] \, [e(u) , f(v) ] = - \frac{p}{2}\frac{ \chi'(u) - \chi'(v) }{u-v}, \\[1mm] \end{array}\\[2mm] \begin{array}{lll} \, [\chi(u) , e(v) ] & = & - \frac{p}{u-v} \{ 2+p \chi'(u), e(u) - e(v) \} \\[1mm] \, [\chi(u) , f(v) ] & = & \frac{p}{u-v} \{ 2+p \chi'(u), f(u) - f(v) \} \\[1mm] \end{array}\\[2mm] \end{array} \end{equation} \begin{equation} \label{lim-cyg} \begin{array}{lcl} \Delta (e(u)) & = & e(u) \otimes 1 + 1 \otimes e(u) + \frac{p}{2} \chi'(u) \otimes e(u), \\[1mm] \Delta (f(u)) & = & f(u) \otimes 1 + 1 \otimes f(u) + \frac{p}{2} f(u) \otimes \chi'(u), \\[1mm] \Delta (\chi'(u)) & = & \chi'(u) \otimes 1 + 1 \otimes \chi'(u) + \frac{p}{2}(\chi'(u) \otimes \chi'(u)), \\[1mm] \Delta (\chi(u)) & = & \chi(u) \otimes 1 + 1 \otimes \chi(u) + \\[1mm] & & +\frac{p}{2}(\chi'(u) \otimes \chi(u) + \chi(u) \otimes \chi'(u)) \\[1mm] & & - 4p f(u) (1 + \frac{p}{2} \chi'(u)) \otimes (1 + \frac{p}{2} \chi'(u)) e(u),\\[1mm] \end{array} \end{equation} where $e(u), f(u)$ and $\chi(u)$ are as in (\ref{gen-fs}) while $\chi'(u)$ is the analog of $\chi(u)$ for the generator $h'$. This Hopf algebra $Y_{p,0}(a)$ is a Yangian for a nonsemisimple Lie algebra $a$ with the following compositions: \begin{equation} \label{lim-g} \begin{array}{c} \, [h'_0, h_0] = 0,\\[1mm] \begin{array}{ll} \, [h_0, e_0] =4pe_0, & [h'_0, e_0] =0, \\[1mm] \, [h_0, f_0] =-4pf_0, & [h'_0, f_0] =0, \\ \end{array}\\[2mm] \, [e_0, f_0] =\frac{p}{2}h'_0, \end{array} \end{equation} This is just the cotangent bundle algebra for the two-dimensional Borel subalgebra of sl(2). For possible applications it is crutially important to know whether the Yangians like $Y_{p,o}(a)$ are pseudotriangular or not, that is whether they have the universal ${\cal R}$-matrix providing the property (\ref{pqybe}). Below we shall present the indications that such ${\cal R}$-matrises exist. The direct way to solve the problem is to construct the parametrized version of the expansion terms for ${\cal R}$-matrix in the general case of $Y_{p,t}(sl(2))$ and to check their limits. In our case this aproach doesn't work. The first nontrivial term -- the classical $r$-matrix after being parametrized according to the transformation (\ref{subs}) diverges in the limit point. As we have seen above the limiting procedure must be accompanied by the rescaling of commutators. This clearly indicates that certain terms of the $r$-matrix must be rescaled in the neighborhood of the limit point. It is not difficult to find these terms and to perform an adequate rescaling. The result is formulated in the statement that follows. \underline{Theorem.} {\sl The boundary Yangian $Y_{p,0}(a)$ ( correspondingly $Y_{0,t}(a))$ originating from $Y(sl(2))$ is the deformation quantization of the polinomial current algebra based on the algebra $a$ (\ref{lim-g}). The first order expansion term for this deformation is defined by the following solution of the classical Yang-Baxter equation for the algebra $a$: \begin{equation} \label{rat-sol2} r(u,v) = \frac{1}{u-v} \left[ \frac{1}{8}(h_0 \otimes h'_0 + h'_0 \otimes h_0) + e_0 \otimes f_0 + f_0 \otimes e_0 \right]. \end{equation} } The first assertion becomes obvious when the classical limit of $Y_{p,0}(a)$ is considered. The validity of the last statement can be checked by the direct computation of the dual Lie algebra defined by (\ref{rat-sol2}); it coinsides with the co-Lie structure that one observes in (\ref{lim-cy}) or (\ref{lim-cyg}). To make the demonstration most transparent we have studied the simpliest possible case -- the Yangian based on $sl(2)$ algebra. As a result our boundary Yangian $Y(a)$ admits some additional simplifications. The subalgebra generated by $\chi'(u)$ forms a Hopf ideal $J(\chi'(u)) \in Y(a)$. In the factor algebra $\frac{Y(a)}{J(\chi'(u))} \equiv Y(c) $ the only nontrivial relations are \begin{equation} \label{fact-m} \begin{array}{lll} \, [\chi(u) , e(v) ] & = & - \frac{4p}{u-v} (e(u) - e(v)), \\[1mm] \, [\chi(u) , f(v) ] & = & \frac{4p}{u-v} (f(u) - f(v)) \end{array} \end{equation} and \begin{equation} \label{fact-c} \Delta (\chi(u)) = \chi(u) \otimes 1 + 1 \otimes \chi(u) - 4p f(u) \otimes e(u). \end{equation} This Yangian $Y(c) \approx U_{\cal F}(c[u])$ is a quantized algebra of currents for the Lie algebra $c$, where $c$ is the algebra $(sl(2))^{\rm contr}$ with the trivially contracted composition $ [e,f]=0 $. The defining relations show that the multiplications in $Y(c)$ are undeformed; the commutators are classical and coinside with those of $U(c [ u ] )$, \begin{equation} \label{fact-m-g} \begin{array}{llllll} \, [h_k,h_l] & = & 0; & [h_0,e_l] & = & 4pe_l ; \\ \, [e_k,f_l] & = & 0; & [h_0,f_l] & = & -4pf_l ; \\ \, [e_{k+1},e_l] & = & [e_{k},e_{l+1}]; & [f_{k+1},f_l] & = & [f_{k},f_{l+1}] ;\\ \, [h_{k+1},e_l] & = & [h_{k},e_{l+1}]; & [h_{k+1},f_l] & = & [h_{k},f_{l+1}] . \end{array} \end{equation} It also has trivial coproducts for $e_k$, $f_k$ \, $(k \in {\bf Z}^+)$ and $h_0$. The nontrivial costructure in $Y(c)$ is generated by the relation $$ \Delta (h_1) = h_1 \otimes 1 + 1 \otimes h_1 - 4pf_0 \otimes e_0 $$ and the compositions (\ref{fact-m-g}). This quantized algebra can be obtained from $U(c[u])$ by a twisting procedure: $U(c[u]) \stackrel{\cal F}{\rightarrow} Y(c)$. The carrier of this twist is an abelian subalgebra generated by the primitive elements $e_0$ and $f_0$ and the twisting element has the form $$ {\cal F}(u-v) = \exp{\left( \frac{1}{v - u} f_0 \otimes e_0 \right)}. $$ This gives the possibility to write down the universal element for $Y(c)$, $$ {\cal R}(u-v) = \exp{ \left( \frac{1}{u-v} ( f_0 \otimes e_0 + e_0 \otimes f_0 ) \right) }. $$ It can be checked that this ${\cal R}$-matrix provides the pseudotriangularity of $Y(c)$. The first nontrivial term of the power series expansion of ${\cal R}$ coinsides with the classical $r$-matrix that can be obtained from the expression (\ref{rat-sol2}) after the factorization by the ideal $J(h'_k)|_{k \in {\bf Z}^{+}}$. \section{Conclusions} We have demonstrated that dual structures signify the existence of the nontrivial limiting algebraic objects corresponding to the rational solutions of CYBE based on nonsemisimple Lie algebras. Thus rational solutions of the type (\ref{rat-sol}) are not necessarily connected with the nondegeneracy of the Killing form. The existence of such rational solutions for the so called symmetric algebras was established in \cite{STOBF}. It was also mentioned there that probably Yangians for symmetric algebras exist. Our result proves this supposition. An important subclass of symmetric algebras is presented by Manin triples. These are the algebras $(a, b(2), b(2)^*)$ (with two-dimensional Borel algebra $b(2)$) that form a Manin triple in our case and its characteristic nondegenerate form was used to construct the invariant element in (\ref{rat-sol2}). The procedure proposed above is quite general and can be applied to any Hopf algebra that have the form of deformation quantization. For example, it can be applied to the Yangian double (DY) where the dual structure is more simple than in the case of the Yangian itself -- the double is not factorized in DY. It must be noted that in this case the situation with the canonical limits for the ${\cal R}$-matrix is complicated by the specific form of dualization established for DY \cite{KHOR}. It depends on the quantization parameter and fails in the canonical limits. To reobtain quasitriangularity for the corresponding boundary Yangian one must reformulate the structures of the Yangian double in terms of the canonical dualization. It is significant that after the factorization the boundary Yangian can be presented in the form of a twist so that its ${\cal R}$-matrix can be written explicitly. {\section*{Acknowledgments}} \noindent One of the authors (L.V.D.) expresses his sincere gratitude to the organizers of the International Workshop SQMQO in Burgos for their hospitality. \bigskip
\section{Introduction} The major problem in the study of the transition to quark matter is the uncertainty in the choice of the relevant degrees of freedom at the densities and temperatures at which the transition is supposed to take place. In the present contribution I will restrict the discussion to the low temperatures regime, {\it i.e.} temperatures of the order of few tens MeV. In this regime, the transition to quark matter will be driven essentially by the density of the system, while the temperature will play a relatively minor role, shifting the critical densities to lower values. The reasons why I will concentrate into the low temperature regime are the following: a) neutron stars (NS), except during the very first seconds of their life, are at temperatures of the order of few MeV or lower; b) during the pre-supernova collapse, temperatures not exceeding few tens MeV can be reached; finally c) quark models have been extensively studied at low temperatures, where the experimental value of several observables constrains the value of the models' parameters. In this contribution I will try to show that, at least in the above described regime, a coherent scenario can be outlined. \section{Quark models} Most of the calculations of the transition to quark matter are based on the MIT bag model. The crucial parameter in this model is the so-called {\it pressure of the vacuum} $B$, namely the energy necessary to dig a hole in the non-perturbative QCD vacuum. Here are a few estimates of $B$: \begin{itemize} \item from the computation of the hadronic spectrum \cite{degrand} $B=59$ MeV/fm$^3$ or $B^{1/4}=145$ MeV \item from the computation of hadronic structure functions \cite{thomas} $B\sim 109$ MeV/fm$^3$ or $B^{1/4} \sim 170$ MeV \item from the comparison of MIT bag pure-gauge results with lattice QCD at finite temperature \cite{satz}: a critical temperature $T_c= 240\pm20$ MeV for pure gauge SU(3) corresponds to $B=170 \pm 50 $ MeV/fm$^3$ or $B^{1/4}= 190\pm 15$ MeV \end{itemize} The conclusions one can obtain from the previous estimates are the following: a) there is a certain ambiguity among the various calculations of {\it single hadron} properties. The preferred value for $B$ is in the range 60 MeV/fm$^3 \le B \le $ 110 MeV/fm$^3$; b) the estimates based on calculations trying to reproduce lattice QCD results indicate a larger value for $B$. It is anyway important to stress that at the moment there is no consensus about the relevant degrees of freedom at finite temperature. For instance, gluons can develop a thermal mass: if the latter is taken into account the lattice QCD data can be reproduced using a very small value for $B$ \cite{greiner}. It is also important to stress that for B$\ge$ 65 MeV/fm$^3$ iron is the ground state of matter. Therefore one is not obliged to accept Witten's strange matter hypothesis even using for B a value indicated by single hadron calculations. Among the many studies of the structure of NS based on the MIT bag model, I would like to comment the most recent ones. Akmal {\it et al.} \cite{akmal} have used, for the hadronic sector, a non-relativistic Equation Of State (EOS) based on the Argonne potential and incorporating three-body forces. In their analysis they consider two values for $B$: 122 and 200 MeV/fm$^3$. Using the smaller value the transition to quark matter starts at $\rho_c\sim 3.5 \rho_0$ and the bulk of a heavy NS is made of a mixed phase of hadronic and quark matter. Clearly, if a smaller value for $B$ would have been used, the mixed phase would be present also in lighter NS. A similar calculation has been performed by IIda and Sato \cite{iida}. In this case in the hadronic sector it has been used a relativistic EOS which takes into account also hyperonic degrees of freedom. The main result they obtain is that heavy NS contain quark matter essentially for all values of $B$. Moreover, if $B$ is smaller than $\sim$ 90 MeV/fm$^3$ quark matter is present even in `standard' NS having a mass $M\sim$1.4 M$_\odot$. The calculations that will be presented in the remaining part of this contribution are not based on the MIT bag model, but on a non-topological soliton model called Color Dielectric Model (CDM) \cite{pirner,birse,banerjee}. The main differences respect to the MIT bag model are the following: \begin{itemize} \item MIT \begin{itemize} \item current masses for the quarks \item large value of the pressure of the vacuum \item rigid confinement in a sphere \item ambiguous treatment of center of mass motion \end{itemize} \item CDM \begin{itemize} \item constituent masses for the quarks \item small value of the pressure of the vacuum \item soft confinement {\it via} interaction with a scalar field \item center of mass motion removed in the non-relativistic limit \end{itemize} \end{itemize} In the past it has been shown that in the CDM, using a fixed set of parameters' values it is possible to study nucleon form factors, structure functions and to reproduce the main features of low energy spectroscopy. This same set of parameters' values has been used to study the transition to quark matter. \section{Neutron stars} It has been investigated the structure of a {\it hybrid} star, which in the outer region is made of nucleonic matter and in the center contains quark matter \cite{plb}. \subsection{Composition} In most of the calculations a naive Walecka-type relativistic model has been used to describe the hadronic phase. Considering a NS having a mass $M=1.4 M_\odot$, half of the volume is occupied by pure quark matter. The mixed phase extends over 1.5 Km. The radius of the star is slightly larger than 10 Km. The maximum mass for a non-rotating star is $M_{max}=1.59 M_\odot$. Looking to the mass-radius relation it appears that the star is more compact than a NS made only of nucleons. This can be relevant in the light of recent estimates of the radii of NS \cite{bombaci}. It is worth mentioning the recently discovered possibility of having, during the slow-down of millisecond pulsars, an effect similar to the back-bending in nuclear physics \cite{glendenning}. Since a millisecond pulsar is strongly deformed, its central density is reduced respect to a non-rotating star. It is therefore possible that quark matter is formed during the slow-down of the pulsar. Since quark matter is denser than normal matter, the moment of inertia of the star reduces and the pulsar re-accelerates to conserve the angular momentum. If detected, this anomalous behaviour of a millisecond pulsar would be the signal that a deep modification in the composition of the star is taking place. \subsection{Cooling} The problem of the cooling of a NS has become very unclear in the last years \cite{page}. It has been pointed out recently \cite{chabrier} that the relation between the internal temperature of the NS, which depends on the composition of the interior of the star, and the external temperature, that can be measured using X-ray satellites, can be deeply modified by the presence of light nuclei on the surface of the star. The presence of light nuclei increases the heat transport and therefore the ratio between the internal and the external temperature could be smaller than the previously estimated factor one hundred. It is anyway interesting to remark that, using the CDM, the direct URCA mechanism at the level of the quarks can take place, but is strongly suppressed due to the extreme smallness of the electron fraction in the interior of the NS. The computed cooling rate is only slightly faster than the modified URCA mechanism taking place in a traditional NS made of nucleons. Therefore, at variance with the MIT bag results, using the CDM is not necessary to invoke any re-heating mechanism to obtain a cooling rate of the right order of magnitude. \section{Supernova explosion} The problem of getting a successful supernova explosion is still open after many years of work. I will concentrate here on the so-called direct mechanism. It is characterized by the idea that the explosion is directly related to the shock wave generated by the bounce. During the last years this mechanism drop out of fashion, and most of the recent studies are related to the idea of a neutrino driven explosion, in which the initial shock stalls and the explosion is revitalized by the energy deposited by the neutrinos in the regions outside the forming proto-neutron star. \begin{figure} \centerline{\hbox{ \psfig{figure=trans.ps,angle=90,height=7truecm,width=10truecm} }} \caption{Boundaries separating hadronic matter from mixed phase (left lines) and the latter from quark matter (right lines). The labels indicate various values of $Y_{l_e}$, $s$ is for symmetric matter. } \end{figure} The possibility of a successful explosion via the direct mechanism is related to the softness of the EOS at densities just above nuclear matter saturation density \cite{swesty}. This possibility has been ruled out for the only reason that seems to be inconsistent with the 1.44 $M_\odot$ constraint coming from PSR 1913+16 \cite{arnett}. On the other hand Cooperstein concluded that if a phase transition takes place during the collapse, `the presence of a mixed-phase region softens the EOS and leads to a direct explosion' \cite{cooperstein}. The crucial problem to be analyzed is the dependence of the critical densities on the proton fraction $Z/A$. Clearly no quark matter has to be present in (nearly) symmetric nuclear matter at densities of the order of the saturation density $\rho_0$. On the other hand, it is exactly in this range of densities that the transition should take place, when $Z/A\sim 0.3$, to influence the collapse of the supernova. We have therefore explored the dependence of the critical densities on the proton fraction \cite{astro}. It is also important to check the effect of the temperature on the transition. In Fig. 1 are presented the boundaries separating hadronic matter from mixed phase and the latter from pure quark matter. The labels correspond to various values of the lepton fraction $Y_{l_e}$. Symmetric nuclear matter is also presented. The transition region depends on the lepton fraction $Y_{l_e}$. In symmetric matter at low temperatures the mixed phase forms at $\rho=0.23$ fm$^{-3}$, therefore no quark matter is present in heavy nuclei. Decreasing the value of $Y_{l_e}$ the phase transition starts at lower densities. At any value of $Y_{l_e}$ the mixed phase extends on a rather limited range of densities and even at zero temperature pure quark matter phase is reached at densities slightly larger than $2 \rho_0$. At higher temperatures the transition starts at lower densities. \begin{figure} \centerline{\hbox{ \psfig{figure=eos.ps,angle=90,height=7truecm,width=10truecm} }} \caption{Pressure (upper box) and adiabatic index (lower box) as function of the density in CDM (solid) and in BCK (dotted). The pressure in Walecka model is also shown (dashed). } \end{figure} To investigate our EOS in connection with the problem of supernova explosion, we compare with BCK EOS \cite{bck}. The latter is a totally phenomenological EOS which is soft enough to allow for supernova explosion, but gives a maximum mass smaller than the mass of PSR 1913+16. In Fig. 2 are presented results for $Y_{l_e}=0.4$ and entropy per baryon number $S/R=1$. In the upper box we compare the pressure in the Walecka model, in our model and in BCK EOS. Due to the phase transition, our EOS is rather soft from $\rho=0.17$ fm$^{-3}$ to $\rho=0.34$ fm$^{-3}$. On the other hand, after $\rho = 0.34$ fm$^{-3}$ it is considerably stiffer than BCK, allowing higher masses for the proto-neutron star. These conclusions are strengthened by the computation of the adiabatic index, shown in the lower box of Fig. 2. Clearly in the mixed phase matter offers little resistance to collapse, but when pure quark matter phase is reached the collapse is halted. In the mixed phase region our adiabatic index is even smaller than in BCK. It is important to remark that similar results have been obtained using the Lattimer-Swesty EOS \cite{ls} to describe the hadronic phase. \section{Conclusions} The main results are the following. Calculations based on various quark models indicate the presence of quark matter at least in the center of heavy NS. Moreover, if the quark model's parameters are fixed to reproduce basic hadronic properties then a) the transition to quark matter takes place at low densities in $\beta$-stable matter and b) neutron stars having a mass $M \sim 1.4 M_\odot$ contain quark matter. Hybrid stars are characterized by a small $R/M$ ratio. A signal of a transition to quark matter would be the spin-up of a millisend pulsar, due to the modification of its moment of inertia. Finally, the transition to quark matter could take place during heavy star collapse, helping supernova to explode. \section*{References}
\section{Introduction} The discovery of ``gaps'' along the blue horizontal branch (HB) in globular clusters as well as of long extensions towards higher temperatures has triggered several spectroscopic investigations (Moehler \cite{moeh99} and references therein) yielding the following results: \begin{enumerate} \item Most of the stars analysed above and below any gaps along the blue horizontal branch are ``bona fide'' blue HB stars ($T_{\rm eff} < 20,000$~K), which show significantly lower gravities than expected from canonical stellar evolution theory. \item Only in NGC~6752 and M~15 have spectroscopic analyses verified the presence of stars that could be identified with the subdwarf B stars known in the field of the Milky Way ($T_{\rm eff} > 20,000$~K, \logg\ $>$ 5). In contrast to the cooler blue HB (BHB) stars the gravities of these ``extended HB'' (EHB) stars agree well with the expectations of canonical stellar evolution. \end{enumerate} Two scenarios have been suggested to account for the low gravities of BHB stars: \begin{description} \item [{\bf Helium mixing scenario:}] Abundance anomalies observed in red giant branch (RGB) stars in globular clusters (e.g., Kraft \cite{kraf94}, Kraft et al. \cite{krsn97}) may be explained by the dredge-up of nuclearly processed material to the stellar surface. If the mixing currents extend into the hydrogen-burning shell -- as suggested by current RGB nucleosynthesis models and observed Al overabundances -- helium can be mixed into the stellar envelope. This in turn would increase the luminosity (and mass loss) along the RGB (Sweigart \cite{swei99}) and thereby create less massive (i.e. bluer) HB stars with helium-enriched hydrogen envelopes. The helium enrichment increases the hydrogen burning rate, leading to higher luminosities (compared to canonical HB stars of the same temperature) and lower gravities. The gravities of stars hotter than about 20,000~K are not affected by this mixing process because these stars have only inert hydrogen shells. \item [{\bf Radiative levitation scenario:}] Grundahl et al. (\cite{grca99}) found a ``jump'' in the $u$, $u-y$ colour-magnitude diagrams of 15 globular clusters, which can be explained if radiative levitation of iron and other heavy elements takes place over the temperature range defined by the ``low-gravity'' BHB stars. This assumption has been confirmed in the case of M~13 by the recent high resolution spectroscopy of Behr et al. (\cite{beco99}). Grundahl et al. argue that super-solar abundances of heavy elements such as iron should lead to changes in model atmospheres which may be capable of explaining the disagreement between models and observations over the ``critical'' temperature range $11,\!500~{\rm K} < T_{\rm eff} < 20,\!000$~K. \end{description} NGC~6752 is an ideal test case for these scenarios, since it is a very close globular cluster with a long blue HB extending to rather hot EHB stars. While previous data already cover the faint end of the EHB, we now obtained new spectra for 32 stars in and above the sparsely populated region between the BHB and the EHB stars. In this {\em Letter}, we present atmospheric parameters derived for a total of 42 BHB and EHB stars and discuss the constraints they may pose on the scenarios described above. \section{Observational Data} We selected our targets from the photographic photometry of Buonanno et al. (\cite{buca86}) to cover the range 14.5$\le V \le$15.5. 19 stars were observed with the ESO 1.52m telescope (61.E-0145, July 22-25, 1998) and the Boller \& Chivens spectrograph using CCD \# 39 and grating \# 33 (65~\AA/mm). This combination covered the 3300~\AA\ -- 5300~\AA\ region at a spectral resolution of 2.6~\AA . The data reduction will be described in Moehler et al. (\cite{mosw99}). Prompted by the suggestion of Grundahl et al. (\cite{grca99}) that radiative levitation of heavy metals may enrich the atmospheres of BHB stars, we searched for metal absorption lines in these spectra. Indeed we found \ion{Fe}{II} absorption lines in almost all spectra (for examples see Fig.~\ref{n6752bhb_speciron}). \begin{figure} \vspace{7.cm} \special{psfile=n6752bhb_speciron.ps hscale=35 vscale=35 hoffset=-20 voffset=200 angle=270} \caption{The iron and magnesium lines as seen in the spectra of some of the stars in NGC~6752. The solid line marks a model spectrum for which we used a solar metallicity model stratification but adjusted all metals to [M/H] = $-$1.5 for the spectrum synthesis except iron. The iron abundance was then adjusted to the noted values to reproduce the marked \ion{Fe}{II} lines. The dashed line marks a model spectrum for which we used the cluster metal abundance for all metals. Effective temperature and surface gravity for both models are those plotted in Fig.~2, upper and central panel, respectively. The temperatures of the stars range from 12,000~K (B3348) to 16,000~K (B1152). Obviously iron is strongly enriched whereas magnesium is consistent with the mean cluster abundance.\label{n6752bhb_speciron}} \end{figure} 13 stars were observed as backup targets at the NTT during observing runs dedicated to other programs (60.E-0145, 61.E-0361). The observations and their reduction are described in Moehler et al. (\cite{mola99}). Those spectra have a spectral resolution of 5~\AA\ covering 3350 to 5250~\AA. No metal lines could be detected due to this rather low spectral resolution. \section{Atmospheric Parameters} \subsection{Fit procedure and model atmospheres} To derive effective temperatures, surface gravities and helium abundances we fitted the observed Balmer lines H$_\beta$ to H$_{10}$ (excluding H$_\epsilon$ because of possible blending problems with the \ion{Ca}{II} H line) and the helium lines (\ion{He}{I} 4026, 4388, 4471, 4922\AA) with stellar model atmospheres. We corrected the spectra for radial velocity shifts, derived from the positions of the Balmer and helium lines and normalized the spectra by eye. We computed model atmospheres using ATLAS9 (Kurucz \cite{kuru91}) and used Lemke's version of the LINFOR\footnote{For a description see http://a400.sternwarte.uni-erlangen.de/$\sim$ai26/linfit/linfor.html} program (developed originally by Holweger, Steffen, and Steenbock at Kiel University) to compute a grid of theoretical spectra which include the Balmer lines H$_\alpha$ to H$_{22}$ and \ion{He}{I} lines. The grid covered the range 7,000~K~$\leq$~\teff~$\leq$~35,000~K, 2.5~$\leq$~\logg~$\leq$~5.0, $-3.0$~$\leq$~\loghe~$\leq$~$-1.0$, at a metallicity of [M/H]~=~$-1.5$. To establish the best fit we used the routines developed by Bergeron et al. (\cite{besa92}) and Saffer et al. (\cite{saff94}), which employ a $\chi^2$ test. The fit program normalizes model spectra {\em and} observed spectra using the same points for the continuum definition. The results are plotted in Fig.~\ref{n6752bhb_irontg} (upper panel). The errors are estimated to be about 10\% in \teff\ and 0.15~dex in \logg\ (cf. Moehler et al. \cite{mohe97}). Representative error bars are shown in Fig.~\ref{n6752bhb_irontg} . To increase our data sample we reanalysed the NTT spectra described and analysed by Moehler et al. (\cite{mohe97}). For a detailed comparison see Moehler et al. (\cite{mosw99}). \begin{figure} \vspace{11.5cm} \special{psfile=n6752bhb_irontg.ps hscale=42 vscale=42 hoffset=-0 voffset=-5 angle=0} \caption{Temperatures and gravities of the programme stars in NGC~6752. {\bf upper panel}: determined from models with cluster metallicity ([M/H] = $-$1.5), {\bf central panel}: adopting a solar metallicity model stratification ([M/H] = 0) and spectrum synthesis with solar iron abundance but cluster abundances for all other metals M ([M/H]=$-$1.5) {\bf lower panel}: adopting a super--solar metallicity model stratification ([M/H] = $+$0.5) and iron abundance ([Fe/H] = $+$0.5) but cluster abundances ([M/H] = $-$1.5) for all other metals in the spectrum synthesis. For more details see text. Also plotted are the zero-age HB (ZAHB) and terminal-age HB (TAHB, i.e., central helium exhaustion) from the Sweigart (1999) tracks for metallicity [M/H] = -1.56. The dashed and solid lines correspond to tracks with and without mixing, respectively. $\Delta X_{\rm mix}$ measures the difference in hydrogen abundance $X$ between the envelope ($X = X_{\rm env}$) and the innermost point reached by the mixing currents ($X = X_{\rm env}-\Delta X_{\rm mix}$) in the red giant precursors and is thus an indicator for the amount of helium mixed into the envelope of the red giant. Representative error bars are plotted\label{n6752bhb_irontg}} \end{figure} \subsection{Iron abundances} Due to the spectral resolution and the weakness of the few observed lines a detailed abundance analysis (such as that of Behr et al., 1999) is beyond the scope of this paper. Nevertheless we can estimate the iron abundance in the stars by fitting the \ion{Fe}{II} lines marked in Fig.~\ref{n6752bhb_speciron}. A first check indicated that the iron abundance was about solar whereas the magnesium abundance was close to the mean cluster abundance. As iron is very important for the temperature stratification of stellar atmospheres we tried to take the increased iron abundance into account: We used ATLAS9 to calculate a solar metallicity atmosphere. The emergent spectrum was then computed from the solar metallicity model stratification by reducing the abundances of all metals M (except iron) to the cluster abundances ([M/H] = $-$1.5). It was not possible to compute an emergent spectrum that was fully consistent with this iron-enriched composition, since the ATLAS9 code requires a scaled solar composition. We next repeated the fit to derive \teff , \logg, and \loghe\ with these enriched model atmospheres. The results are plotted in Fig.~\ref{n6752bhb_irontg} (central panel). \begin{figure} \vspace{12cm} \special{psfile=n6752bhb_iron.ps hscale=45 vscale=45 hoffset=-12 voffset=-10 angle=0} \caption{The iron and helium abundances for the stars observed with ESO 1.52m telescope. Iron was not detected in the coolest star and is plotted as an upper limit. The trend to lower helium abundances for higher temperatures agrees with the findings of Behr et al. (1999). Iron is obviously enhanced to roughly solar abundances. The mean iron abundance as derived from our spectra ([Fe/H] = $+$0.13) and the cluster abundance ([Fe/H] = $-$1.54) are marked. The asterisk marks the results of Glaspey et al. (1989) for the hotter of their two BHB stars in NGC~6752\label{n6752bhb_iron}} \end{figure} For each star observed at the ESO 1.52m telescope we then computed an ``iron-enriched'' model spectrum with \teff , \logg\ as derived from the fits of the Balmer and helium lines with the ``enriched'' model atmospheres (cf. Fig.~\ref{n6752bhb_irontg}, central panel) and \loghe\ = $-$2. The fit of the iron lines was started with a solar iron abundance and the iron abundance was varied until $\chi^2$ achieved a minimum. As the radiative levitation in BHB stars is due to diffusion processes (which is also indicated by the helium deficiency found in these stars) the atmospheres have to be very stable. We therefore kept the microturbulent velocity $\xi$ at 0\,km/s -- the iron abundances plotted in Fig.~\ref{n6752bhb_iron} are thus upper limits. The mean iron abundance turns out to be [Fe/H] $\approx +0.1$~dex (for 18 stars hotter than about 11,500~K) and $\le-$1.6 for the one star cooler than 11,500~K. Although the iron abundance for the hotter BHB stars is about a factor of 50 larger than the cluster abundance, it is smaller by a factor of 3 than the value of [Fe/H] = $+$0.5 estimated by Grundahl et al. (\cite{grca99}) as being necessary to explain the Str\"omgren $u$-jump observed in $u$, $u-y$ colour-magnitude diagrams. Our results are in good agreement with the findings of Behr et al. (\cite{beco99}) for BHB stars in M~13 and Glaspey et al. (\cite{glmi89}) for two BHB stars in NGC~6752. Again in agreement with Behr et al. (\cite{beco99}) we see a decrease in helium abundance with increasing temperature, whereas the iron abundance stays roughly constant over the observed temperature range. \subsection{Influence of iron enrichment} From Fig.~\ref{n6752bhb_irontg} it is clear that the use of enriched model atmospheres moves most stars closer to the zero-age horizontal branch (ZAHB). The three stars between 10,000~K and 12,000~K, however, fall {\em below} the canonical ZAHB when fitted with enriched model atmospheres. This is plausible as the radiative levitation is supposed to start around 11,500~K (Grundahl et al. \cite{grca99}) and the cooler stars therefore should have metal-poor atmospheres (see also Fig.~\ref{n6752bhb_iron} where the coolest analysed star shows no evidence of iron enrichment). We repeated the experiment by increasing the iron abundance to [Fe/H]=$+$0.5 (see Fig.~\ref{n6752bhb_irontg} lower panel), which did not change the resulting values for \teff\ and \logg\ significantly. Since HB stars at these temperatures spend most of their lifetime close to the ZAHB, one would expect the majority of the stars to scatter (within the observational error limits) around the ZAHB line in the \logt , \logg--diagram. However, this is not the case for the canonical ZAHB (solid lines in Fig.~\ref{n6752bhb_irontg}) even with the use of iron-enriched model atmospheres (central and lower panels in Fig.~\ref{n6752bhb_irontg}). The scatter instead seems more consistent with the ZAHB for moderate helium mixing (dashed lines in Fig.~\ref{n6752bhb_irontg}). Thus the physical parameters of HB stars hotter than $\approx 11,\!500$~K in NGC~6752, as derived in this paper, are best explained by a combination of helium mixing and radiative levitation effects. \section{Conclusions} Our conclusions can be summarized as follows: \begin{enumerate} \item We have obtained new optical spectra of 32 hot HB stars in NGC 6752 with 11,000 K $<$ \teff $<$ 25,000. When these spectra (together with older spectra of hotter stars) are analysed using model atmospheres with the cluster metallicity ([Fe/H] = $-$1.5), they show the same ``low-gravity'' anomaly with respect to canonical HB models, that has been observed in several other clusters (Moehler 1999). \item For 18 stars with \teff\ $>$ 11,500 K, we estimate a mean iron abundance of [Fe/H] $\approx +$0.1, whereas magnesium is consistent with the cluster metallicity. The hot HB stars in NGC 6752 thus show an abundance pattern similar to that observed in M~13 (Behr et al. \cite{beco99}), which presumably arises from radiative levitation of iron (Grundahl et al. 1999). \item When the hot HB stars are analysed using model atmospheres with an appropriately high iron abundance, the size of the gravity anomaly with respect to canonical HB models is significantly reduced. Whether the remaining differences between observations and canonical theory can be attributed to levitation effects on elements other than iron remains to be investigated by detailed modeling of the diffusion processes in the stellar atmospheres. With presently available models, the derived gravities for HB stars hotter than $\approx 11,\!500$~K are best fit by non-canonical HB models which include deep mixing of helium on the RGB (Sweigart \cite{swei99}). \end{enumerate} \acknowledgements We thank the staff of the ESO La Silla observatory for their support during our observations. S.M. acknowledges financial support from the DARA under grant 50~OR~96029-ZA. M.C. was supported by NASA through Hubble Fellowship grant HF--01105.01--98A awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA under contract NAS~5-26555. We are grateful to the referee, Dr. R. Kraft, for his speedy report and valuable remarks.
\section*{Introduction} In quantum mechanics a mean field theory means that the particle density $\rho(x) = \psi^*(x)\psi(x)$ (in second quantization) tends to a {\it c}--number in a suitable scaling limit. Of course, $\rho(x)$ is only an operator valued distribution and the smeared densities $\rho_f = \int dx\,\rho(x)f(x)$ are (at best) unbounded operators, so norm convergence is not possible. The best one can hope for is strong resolvent convergence in a representation where the macroscopic density is built in. The BCS--theory of superconductivity is of a different type where pairs of creation operators with opposite momentum $\tilde\psi^*(k)\,\tilde\psi^*(-k)$ ($\tilde\psi$ the Fourier transform and with the same provisio) tend to {\it c}--numbers. This requires different types of correlations and one might think that the two possibilities are mutually exclusive. We shall show that this is not so by constructing a pair potential where both phenomena occure simultaneously. On purpose we shall use only one type of fermions as one might think that the spin--up electrons have one type of correlation and the spin--down the other. Also the state which carries both correlations is not an artificial construction but it is the KMS--state of the corresponding Bogoliubov Hamiltonian. Whether the phenomenon occurs or not depends on whether the emerging two coupled ``gap equations" have a solution or not. This will happen to be the case in certain regions of the parameter space (temperature, chemical potential, relative values of the two coupling constants). For simple forms of the potentials these regimes will be explicitly shown. Our considerations hold for arbitrary space dimension. \section{Quadratic fluctuations in a KMS--state} The solvability of the BCS--model \cite{BCS} rests upon the observation \cite{H} that in an irreducible representation the space average of a quasi--local quantity is a {\it c}--number and is equal to its ground state expectation value. This allows one to replace the model Hamiltonian by an equivalent approximating one \cite{NNB}. Remember that two Hamiltonians are considered to be equivalent when they lead to the same time evolution of the local observables \cite{TW}. The same property holds on also in a temperature state (the KMS--state) and under conditions to be specified later it makes the co--existence of other types of phases possible. To make this apparent, consider the approximating (Bogoliubov) Hamiltonian \begin{eqnarray} H_B' &=& \int dp\,\left\{\omega(p)a^*(p)a(p) + \frac{1}{2}\Delta_B(p)\left[a^*(p)a^*(-p) + a(-p)a(p)\right]\right\} \nonumber \\ &=& \int W(p) b^*(p)b(p)\, , \end{eqnarray} which has been diagonalized by means of a standard Bogoliubov transformation with real coefficients (the irrelevant infinite constant in $H_B'$ has been omitted) \begin{eqnarray*} b(p) = c(p)a(p) + s(p)a^*(-p) \\ a(p) = c(p)b(p) - s(p)b^*(-p) \end{eqnarray*} with $$ c(p) = c(-p) \qquad s(p) = -s(-p) $$ \begin{equation} c^2(p) + s^2(p) = 1\, , \end{equation} so that the following relations hold (keeping in mind that $\Delta, W, s, c$ will be $\beta$--dependent) $$ W(p) = \sqrt{\omega^2(p) + \Delta_B^2(p)} = W(-p) $$ \begin{equation} c^2(p) - s^2(p) = \omega(p)/W(p)\,, \qquad 2c(p)s(p) = \Delta_B(p)/W(p) \end{equation} Hamiltonian (1.1) generates a well defined time evolution and a KMS--state for the $b$--operators. For the original creation and annihilation operators $a, a^*$ this gives the following evolution $$ a(p) \rightarrow a(p)\left(c^2(p)e^{-iW(p)t} + s^2(p)e^{iW(p)t}\right) - 2ia^*(-p)c(p)s(p)\sin W(p)t $$ and nonvanishing termal expectations \begin{eqnarray} \langle a^*(p)a(p')\rangle &=& \delta(p-p')\left\{\frac{c^2(p)}{1+e^{\beta(W(p)-\mu)}} + \frac{s^2(p)}{1+e^{-\beta(W(p)-\mu)}}\right\} \nonumber \\[2pt] &:=& \delta(p-p')\{p\}\\ \langle a(p)a(-p')\rangle &=& \delta(p-p')c(p)s(p)\tanh\frac{\beta(W(p)-\mu)}{2} := \delta(p-p')[p] \end{eqnarray} $$ \{p\} = \{-p\}, \qquad [p] = -[-p] $$ $c$ and $s$ are multiplication operators and are never Hilbert--Schmidt. Thus different $c$ and $s$ lead to inequivalent representations and should be considered as different phases of the system. The expectation value of a biquadratic (in creation and annihilation operators) quantity is expressed through (1.4,5) \begin{eqnarray} \langle a^*(q)a^*(q')a(p)a(p')\rangle = \delta(q+q')\delta(p+p')[q][p]-& \nonumber \\ - \delta(p-q)\delta(p'-q')\{p\}\{p'\} + \delta(p-q')\delta(p'-q)\{p\}\{p'\} & \end{eqnarray} So far we have written everything in terms of the operator valued distributions $a(p)$. They can be easily converted into operators in the Hilbert space generated by the KMS--state by smearing with suitable test functions. Thus, by smearing with e.g. \begin{equation} e^{-\kappa (p+p')^2-\kappa (q+q')^2}v(p)v(q), \qquad v \in L_2({\bf R}^d) \end{equation} one observes that in the limit $\kappa \rightarrow \infty$ the first term in (1.6) remains finite $$ 0 < \int dpdqv(p)v(q)[p][q] < \infty\, , $$ while the two others vanish $$ \lim_{\kappa\rightarrow\infty} \int dpdp' e^{-2\kappa(p+p')^2}v(p)v(p')\{p\}\{p'\} = \lim_{\kappa\rightarrow\infty} \kappa^{-3/2}\int dpv^2(p)\{p\}^2 = 0. $$ Since we are in the situation of {\it Lemma 1} in \cite{IT}, we have thus proved the following statement \begin{equation} \mbox{s-}\lim_{\kappa\rightarrow\infty}\int dpdp'{\cal V}(q,q',p,p')e^{-\kappa(p+p')^2}a(p)a(p') = \int dp{\cal V}(q,q',p,-p)[p] \end{equation} for kernels ${\cal V}$ such that the integrals are finite. With this observation in mind, a potential which acts for $\kappa \rightarrow \infty$ like (1.1) might be written as \begin{equation} V_B = \kappa^{3/2} \int dpdp'dqdq'\,a^*(q)a^*(q')a(p)a(p'){\cal V}_B(q,q',p,p')\, e^{-\kappa(p+p')^2-\kappa(q+q')^2} \end{equation} with ${\cal V}_B(q,q',p,p') = -{\cal V}_B(q',q,p,p')$ etc., in order to respect the fermi nature of $a$'s. This potential has the property $$ \begin{array}{ll} \Vert V \Vert < \infty & \qquad \mbox{ for }\, \kappa <\infty \nonumber \\[4pt] \Vert V \Vert \rightarrow \infty & \qquad \mbox{ for }\, \kappa \rightarrow \infty \nonumber \end{array} $$ Despite this divergence, potential (1.9) may still generate a well--defined time evolution. The strong resolvent convergence in (1.8) is essential, weak convergence would not be enough since it does not guarantee the automorphism property $$ \tau_\kappa^t(ab) = \tau_\kappa^t(a)\tau_\kappa^t(b)\,\rightarrow\, \tau_\infty^t(ab) = \tau_\infty^t(a)\tau_\infty^t(b)\,. $$ Note that the parameter $\kappa$ plays in this construction the role of the volume from the considerations in \cite{H}. In the mean--field regime we want an effective Hamiltonian \begin{equation} H_B'' = \int dp \left[ \omega(p)a^*(p)a(p) + \Delta_M(p)a^*(p)a(p)\right]\, . \end{equation} Here the KMS--state is defined for the operators $a, a^*$ themselves and one should rather smear by means of \begin{equation} e^{-\kappa(q-p)^2-\kappa(q'-p')^2}v(p)v(p') \end{equation} instead of (1.7), thus concluding that \begin{equation} \mbox{ s-}\lim_{\kappa\rightarrow\infty} \int dpdq e^{-\kappa(q-p)^2} a^*(q)a(p) {\cal V}_M(q,q',p,p') = -\int dp\frac{{\cal V}_M(p,q',p,p')}{1 + e^{\beta(\varepsilon(p)-\mu)}}\, , \end{equation} with $\varepsilon(p) = \omega(p) + \Delta_M(p)$. Relation (1.12) then suggests another starting potential \begin{equation} V_M = \kappa^{3/2} \int dpdp'dqdq'\,a^*(q)a^*(q')a(p)a(p'){\cal V}_M(q,q',p,p')\, e^{-\kappa(q-p)^2-\kappa(q'-p')^2} \end{equation} with the same symmetry for the density ${\cal V}_M$ as in (1.9). However, in both cases a Gaussian form factor in the smearing functions (1.7),(1.11) has been chosen just for simplicity. In principle, this might be $C_o^\infty$ functions which have the $\delta$--function as a limit. \section{The model} Consider the following Hamiltonian \begin{equation} H = H_{kin} + V_B + V_M \, , \end{equation} where $H_{kin}$ is the kinetic term and $ V_B, V_M$ are given by (1.9),(1.13). The first potential term describes the superconducting phase, similarly to the BCS--model, while the second corresponds to the mean field regime. As already mentioned, both these terms diverge in the limit $\kappa\rightarrow\infty$. The solvability of the model for $\kappa \rightarrow \infty$ depends on whether or not it would be possible to replace (2.1) by an equivalent Hamiltonian that might be readily diagonalized. Remind that by equivalence of two Hamiltonians an equivalence of the time evolution of the local observables they generate should be understood. Therefore, the object of interest is the commutator of, say, a creation operator with the potential. With (1.8), (1.12) taken into account, it reads \begin{equation} [a(k), V] = 2\int dp\left\{c(p)s(p)\,[p]\,{\cal V}_B(k,-k,p,-p)a^*(-k) + {\cal V}_M(p,k,p,k)\,\{p\}\,a(k)\right\} \end{equation} The Bogoliubov--type Hamiltonian for our problem should be a combination of (1.1) and (1.10), that is of the form \begin{equation} H_B = \int dp \left\{ a^*(p)a(p)[\omega(p) + \Delta_M(p)] + \frac{1}{2}\Delta_B(p)[a^*(p)a^*(-p) + a(-p)a(p)]\right\} \end{equation} This Hamiltonian becomes equivalent to the model Hamiltonian (2.1), provided the commutator $[a(k), H_B - H_{kin}]$ equals (2.2). Thus we are led to a system of two coupled ``gap equations" \begin{eqnarray} \frac{1}{2} \Delta_M(k) &=& \int {\cal V}_M(k,p)\, \left\{\frac{c^2(p)}{1 + e^{\beta(\overline W(p)-\mu)}} + \frac{s^2(p)}{1+e^{-\beta(\overline W(p)-\mu)}}\right\} \,dp \\[10pt] \Delta_B(k) &=& \int {\cal V}_B(k,p)\, \frac{\Delta_B(p)}{\overline W(p)} \,\tanh\frac{\beta(\overline W(p)-\mu)}{2}\,dp\, , \end{eqnarray} with \begin{equation} \overline W(p) = \sqrt{[\omega(p) + \Delta_M(p)]^2 + \Delta_B^2(p)}\, . \end{equation} $c$ (and thus $s$, eq.(1.2)) are determined by either of the following conditions \begin{equation} c^2(p) - s^2(p) = [\omega(p)+\Delta_M(p)]/\overline W(p)\,, \qquad 2c(p)s(p) = \Delta_B(p)/\overline W(p)\, . \end{equation} The temperature and the interaction--strenght dependence of the system (2.4--7) encode the solvability of the model. \section{Solution to the coupled gap equations} For general potential densities ${\cal V}_M, {\cal V}_B$ solutions of the system (2.4--6) cannot be explicitly written. In both low-- and high--temperature limits a non--trivial ``mean--field gap" is possible, while for the superconducting phase, characterized by a non--vanishing $\Delta_B$, a critical temperature exists beyond which such a phase is no longer possible. However, for some simple though reasonable potentials solvability of (2.4-7) and the behaviour of the solutions can be discussed in more detail. Thus, for the special case of an interaction concentrated about the Fermi surface and being constant therein, potential densities can be chosen as $$ {\cal V}_{B,M}({\bf k,p}) = \lambda_{B,M} {\cal S}(\bf k) {\cal S}(\bf p) $$ with $$ {\cal S}({\bf k}) = \frac{1}{2\varepsilon}\, [\Theta(\vert{\bf k}\vert-\sqrt{\mu}+\varepsilon) - \Theta(\vert{\bf k}\vert-\sqrt{\mu}-\varepsilon)], $$ where $\Theta(x)$ is the Heaviside function. With the additional assumption $\omega(p)=p^2$, the system (2.4--6) transforms for $\varepsilon \rightarrow 0$ into \begin{eqnarray} \frac{1}{2} \Delta_M(\mu) &=& \lambda_M\,\left\{\frac{c^2(\mu)}{1 + e^{\beta(\overline W(\sqrt{\mu})-\mu)}} + \frac{s^2(\mu)}{1+e^{-\beta(\overline W(\sqrt{\mu})-\mu)}}\right\} \\[6pt] \overline W(\sqrt{\mu}) &=& \lambda_B\, \tanh\frac{\beta(\overline W(\sqrt{\mu})-\mu)}{2} \qquad \mbox{or }\qquad \Delta_B = 0 \end{eqnarray} \begin{equation} \overline W(\sqrt{\mu}) = \sqrt{[\mu + \Delta_M(\mu)]^2 + \Delta_B^2(\mu)}\\[4pt] \end{equation} with subsidiary conditions (2.7) correspondingly modified. We shall always assume $\overline W(\sqrt{\mu})>0$. This is not really a restriction, since the opposite situation might be similarly treated after performing the exchange $b^* \leftrightarrow b$, also we might take $\mu + \Delta_M \geq 0$. There are four energies $\lambda_M, \lambda_B, \mu, \beta=1/T$ involved. The system exhibits severe dependence on their relative values, in particular, the following holds: \begin{enumerate} \item[{\rm (i)}] for all values of $\lambda_M, \lambda_B, \mu$ there is a purely mean-field solution $$ \Delta_B = 0, \qquad \overline W(\sqrt{\mu}) - \mu = \Delta_M = \frac{2\lambda_M}{1+e^{\beta \Delta_M}}\, , $$ but this will not be considered further; \item[{\rm (ii)}] $\overline W(\sqrt{\mu}) \leq \vert \lambda_B \vert$. Furthermore, the condition $\overline W(\sqrt{\mu})>0$ implies $$ \lambda_B > 0\,\Longleftrightarrow \,\overline W(\sqrt{\mu}) > \mu, \qquad \lambda_B < 0\,\Longleftrightarrow \,\overline W(\sqrt{\mu}) < \mu; $$ \item[{\rm (iii)}] eq.(3.1) tells us that $\rm sign\,\Delta_M = \rm sign\,\lambda_M\,$ and $\,\vert \Delta_M \vert \leq 2\vert \lambda_M \vert\,$; \item[{\rm (iv)}] $\overline W > \mu + \Delta_M\,$ and $\,\overline W \geq \Delta_B$. In addition, eq.(2.7) brings a restriction on the mixing angles in the Bogoliubov transformation, $\varphi \in [-\pi/4,\pi/4]\, \cup\, [3\pi/4,-3\pi/4]$, so that $\sqrt{2}/2~\leq~\vert c(p)\vert \leq 1$. \end{enumerate} Thus nontrivial solutions are only possible in the following regions: \begin{enumerate} \item[{\bf (a)}] $\lambda_B > 0: \quad \lambda_B > \mu\,$ \,\,(the area $\mbox{{{}\hspace{.125em}\raisebox{.32ex}[1.2ex][0ex]{\"{}}\hspace{-.625em}}A}^+\cup {\cal B}^+\cup{\cal C}^+$ on Fig. 5); \item[{\bf (b)}] $\lambda_B < 0: \quad -2\mu \leq \lambda_M \leq -\displaystyle \frac{\mu T}{\vert \lambda_B \vert + 2T}$\\[4pt] (limitations for the area $\mbox{{{}\hspace{.125em}\raisebox{.32ex}[1.2ex][0ex]{\"{}}\hspace{-.625em}}A}^-\cup{\cal B}^-$ on Fig. 5). \end{enumerate} Therefore two general cases have to be distinguished, corresponding to attractive or repulsive superconducting potential, $\lambda_B<0$ or $\lambda_B>0$, that are qualitatively represented for increasing values of $\vert \lambda_B \vert$ on Fig. 1. \begin{figure}[h] \setlength{\unitlength}{1pt} \begin{picture}(490,155) \put(25,0){\makebox(0,0)[lb] {\scalebox{0.9}% {\includegraphics*[0,235][490,410]{bcsf1.eps}}}} \put(200,48){\makebox(0,0)[l]{$x$}} \put(414,48){\makebox(0,0)[l]{$x$}} \end{picture} \caption{$\,\lambda_B\tanh{(x \mp k\lambda_B)}\,$ compared to the straight line for increasing values of $\vert \lambda_B \vert$ and $k = 0.8$; $\,x = \beta \overline W(\sqrt{\mu})/2$: {\it(i)} $\,\lambda_B > 0$; {\it (ii)} $\,\lambda_B < 0$.} \end{figure} \vspace{0.6cm} {\bf I.} \underline{$\lambda_B > 0$} \vspace{6pt} In this case necessarily only excitations with energies $> \mu$ may be present. In both limits $\overline W \gg \mu + 2T$ and $\overline W \ll \mu + 2T$ solutions can be obtained analytically. \begin{itemize} \item[{\bf I.A}]\, $0 < \overline W(\sqrt{\mu}) - \mu \gg 2T, \, \tanh{(x)} \rightarrow 1$ \end{itemize} In this regime one finds \begin{eqnarray} \overline W &=& \lambda_B \\ \Delta_M &=& \lambda_M \frac{\lambda_B - \mu}{\lambda_B + \lambda_M}\\ \Delta_B &=& \pm\frac{\lambda_B}{\lambda_B + \lambda_M}\sqrt{(\lambda_B-\mu)(\lambda_B + \mu + 2\lambda_M)} \end{eqnarray} Thus a restriction has to be fulfilled \begin{equation} \lambda_B + \mu + 2\lambda_M > 0\, , \end{equation} which is always the case by $\lambda_M > 0$ and also by negative $\lambda_M$, if in addition $$ \vert \lambda_M \vert < (\lambda_B + \mu)/2\, , $$ that determines the area $\mbox{{{}\hspace{.125em}\raisebox{.32ex}[1.2ex][0ex]{\"{}}\hspace{-.625em}}A}^+\cup{\cal B}^+$ on Fig. 5. Then also the mean--field gap energy becomes negative, though, as already mentioned, it is possible to demand positivity for the ``pure" mean--field energy, $\mu + \Delta_M > 0$. \vspace{2pt} \begin{itemize} \item[{\bf I.B}]\, $0 < \overline W(\sqrt{\mu}) - \mu \ll 2T, \, \tanh{(x)} \rightarrow x $ \end{itemize} Here the solution reads \begin{eqnarray} \overline W(\mu) &=& \frac{\lambda_B\mu}{\lambda_B - 2T}\\[2pt] \Delta_M &=& \lambda_M \\[2pt] \Delta_B &=& \pm \sqrt{\frac{\lambda_B^2\mu^2}{(\lambda_B-2T)^2} - (\mu+\lambda_M)^2} \end{eqnarray} and exists in the temperature interval \begin{equation} \frac{\lambda_M\lambda_B}{2(\mu+\lambda_M)} < T \ll \frac{\lambda_B-\mu}{2}\, . \end{equation} The above restriction substanciates the idea of this limit as being valid at sufficiently high, but nevertheless not at extremely high temperatures. With an accuracy of $10^{-5}$ one can then estimate $$ \lambda_M < \frac{\lambda_B - 4\mu}{4}\, , $$ so this asymptotic regime, when admissible, corresponds to the area ${\cal B}^+\cup{\cal C}^+$ on Fig. 5. In the intermediate region an interesting phenomenon occurs that might be visualized by the plot of the two sides of eq.(3.2) --- Fig. 2. There are three different sets of values of the parameters of the theory, for which one of the following situations is realised (see also Fig. 1(i)): \begin{figure}[h] \setlength{\unitlength}{1pt} \begin{picture}(470,150) \put(0,0){\makebox(0,0)[lb] {\scalebox{0.9}% {\includegraphics*[0,0][470,150]{bcsf2.eps}}}} \put(140,15){\makebox(0,0)[l]{{\small $\frac{\beta \overline W(\sqrt{\mu})}{2}$}}} \put(41,15){\makebox(0,0)[l]{{\small $\bar \lambda_B$}}} \put(370,15){\makebox(0,0)[l]{{\small $\frac{\beta \overline W(\sqrt{\mu})}{2}$}}} \put(268,15){\makebox(0,0)[l]{{\small $\bar \lambda_B$}}} \end{picture} \caption{Plot of both sides of eq.(3.2) for $\lambda_B > 0\,$: {\it (i)} $\, \mu=0.2\,\bar\lambda_B$; {\it (ii)} $\,\mu=0.9\,\bar\lambda_B$.} \end{figure} \begin{enumerate} \item[(a)] the co--existence of the mean--field regime and the superconducting phase is not possible ($\Delta_M\not=0, \, \Delta_B=0$); \item[(b)] for fixed values of the parameters such a mixed phase is brought in to being and is uniquely determined ($\Delta_M\not=0,\,\Delta_B\not=0$); \item[(c)] a kind of bifurcation occurs, so that two mixed phases with different $\Delta_M$ and $\Delta_B$ become possible. \end{enumerate} \begin{figure}[ht] \setlength{\unitlength}{1pt} \begin{picture}(425,160) \put(12,0){\makebox(0,0)[lb] {\includegraphics*[5,220][430,380]{bcsf3.eps}}} \put(0,127){\makebox(0,0)[l]{{\small $\frac{\beta \overline W}{2}$}}} \put(200,6){\makebox(0,0)[l]{{\small $\bar\lambda_B$}}} \put(240,127){\makebox(0,0)[l]{{\small $\frac{\mu}{\overline W}$}}} \put(432,6){\makebox(0,0)[l]{{\small $\bar\lambda_B$}}} \put(340,65){\makebox(0,0)[l]{{\small (a)}}} \put(425,100){\makebox(0,0)[l]{{\small (b)}}} \put(280,95){\makebox(0,0)[l]{{\small (c)}}} \end{picture} \caption{{\it (i)\/}``Thermalized" energy $\beta\overline W(\sqrt{\mu})/2$ for the equilibrium chemical potential, compared to the spectra for potentials $\mu=k\,\lambda_B,\,\,k=0.1,0.3,0.5,0.7,0.9 \,\,$ ; {\it (ii)\/} Phase diagram for the ratio $\mu/\overline W(\sqrt{\mu})$.} \end{figure} More precisely, there exists a one--parameter family of equilibrium chemical potentials, \begin{equation} \bar\mu_e = -\rm Arccosh{\sqrt{\bar\lambda_B}} + \bar\lambda_B\tanh{(\rm Arccosh{\sqrt{\bar\lambda_B}}\,)}\, , \end{equation} $$ \bar Q := \beta Q/2\, , $$ for which solution of (3.1--3) is uniquely determined. The corresponding energy is then found to be $$ \frac{\beta \overline W(\sqrt{\mu})}{2} = \bar\lambda_B \tanh{(\rm Arccosh{\sqrt{\bar\lambda_B}}\,)}\, . $$ For chemical potentials $\bar\mu < \bar\mu_e$ there is no solution at all, while for $\bar\mu > \bar\mu_e$ there are two solutions. The same is true also for the ratio $\mu/\overline W(\sqrt{\mu})$ --- areas (a), (b) and (c) on Fig. 3(ii) respectively. This, together with the comparison of the ``equilibrium" energy to the spectra for different allowed values of the chemical potential --- Fig. 3(i), shows that for a given value of the coupling constant $\lambda_B$, with the change of the temperature the system can pass from one phase to another, so that a phase transition occurs. When the critical value $T_c = \lambda_B/2$ is then reached, the superconducting phase (either one, or two co--existing such phases) is destroyed again. Furthermore, the two co--existing solutions correspond to two different mixing angles, so are obtained through two different, hence unequivalent, Bogoliubov transformations and this does not directly afflict the local stability of the physical system. \vspace{0.6cm} {\bf II.} \underline{$\lambda_B < 0$} \vspace{6pt} According to (3.2), for negative values of the coupling constant $\,\lambda_B\,$ the excitations are necessarily with energies $\,\overline W < \mu$. Thus, the mean--field gap must be negative, $\,\Delta_M < 0$, that presupposes solvability of the model only when the second coupling constant $\,\lambda_M\,$ is also negative. As one sees on the plot in Fig. 4, there is no nontrivial phase structure present in this regime and the solution is always uniquely determined (see Fig. 1(ii)). \begin{figure}[h] \setlength{\unitlength}{1pt} \begin{picture}(490,190) \put(0,0){\makebox(0,0)[lb] {\scalebox{0.9}% {\includegraphics*[0,300][490,490]{bcsf4.eps}}}} \put(49,35){\makebox(0,0)[l]{{\small $\frac{\beta \overline W}{2}$}}} \put(170,55){\makebox(0,0)[l]{{\small $\bar\lambda_B$}}} \put(280,35){\makebox(0,0)[l]{{\small $\frac{\beta \overline W}{2}$}}} \put(405,55){\makebox(0,0)[l]{{\small $\bar\lambda_B$}}} \end{picture} \caption{Plot of both sides of eq.(3.2) for $\lambda_B < 0$: {\it (i)} $\,\mu < \lambda_B\,$; {\it (ii)} $\,\mu > \lambda_B$.} \end{figure} \vspace{2pt} \begin{itemize} \item[{\bf II.A}]\, $0 < \overline W(\sqrt{\mu}) \ll \mu - 2T $ \end{itemize} The solution looks like the one for positive $\lambda_B$, eqs.(3.4--6), only $\overline W = \vert \lambda_B \vert$. However, the relation between the parameters changes \begin{equation} \lambda_B + \mu + 2\lambda_M < 0 \end{equation} With the assumption $\mu + \Delta_M > 0$ this also means $ \lambda_B < -\mu -\Delta_M $ but this does not strengthen the general restriction {\bf (b)}. \vspace{2pt} \begin{itemize} \item[{\bf II.B}]\, $0 < \overline W(\sqrt{\mu}) \gg \mu - 2T $ \end{itemize} In this case the model is solvable, eqs.(3.8--10), within the temperature interval \begin{equation} \frac{\mu + \lambda_B}{2} \ll T < \frac{\lambda_B\lambda_M)}{2(\mu + \lambda_M)}\, \end{equation} and under the same assumption as in the case $\lambda_B > 0$, one concludes that $$ \lambda_M < -\frac{\lambda_B + 4\mu}{4} \qquad \mbox{ (area }\, {\cal B}^- \, \mbox{on \, Fig.\,5)} $$ The above restrictions for the allowed regions in the $(\lambda_B, \lambda_M)$--plane, are depicted on Fig. 5. Nontrivial solutions are possible in the area $\mbox{{{}\hspace{.125em}\raisebox{.32ex}[1.2ex][0ex]{\"{}}\hspace{-.625em}}A}^\pm\cup{\cal B}^\pm\cup{\cal C}^+$. In the $\mbox{{{}\hspace{.125em}\raisebox{.32ex}[1.2ex][0ex]{\"{}}\hspace{-.625em}}A}^\pm$--regions in the parameter space only a highly excited system exhibits a superconducting behaviour, while in the $\,{\cal B}^\pm$--regions this becomes possible also for a system whose energy is close to the chemical potential and in a restricted temperature interval. In both cases, when all parameters fixed, solutions are uniquely defined. \begin{figure}[ht] \setlength{\unitlength}{1pt} \begin{picture}(320,210) \put(60,0){\makebox(0,0)[lb] {\includegraphics*[0,340][320,550]{bcsf5.eps}}} \put(178,107){\makebox(0,0)[l]{{\small $\mbox{{{}\hspace{.125em}\raisebox{.32ex}[1.2ex][0ex]{\"{}}\hspace{-.625em}}A}^-$}}} \put(140,90){\makebox(0,0)[l]{{\small ${\cal B}^-$}}} \put(208,51){\makebox(0,0)[l]{{\scriptsize -$2\mu$}}} \put(208,85){\makebox(0,0)[l]{{\footnotesize -$\mu$}}} \put(208,103){\makebox(0,0)[l]{{\footnotesize -$\frac{\mu}{2}$}}} \put(180,200){\makebox(0,0)[l]{{\small $\lambda_M$}}} \put(225,129){\makebox(0,0)[l]{{\small $\mu$}}} \put(262,156){\makebox(0,0)[l]{{\small $\mbox{{{}\hspace{.125em}\raisebox{.32ex}[1.2ex][0ex]{\"{}}\hspace{-.625em}}A}^+$}}} \put(294,85){\makebox(0,0)[l]{{\small ${\cal B}^+$}}} \put(250,30){\makebox(0,0)[l]{{\small ${\cal C}^+$}}} \put(350,115){\makebox(0,0)[l]{{\small $\lambda_B$}}} \end{picture} \caption{Allowed regions in the $(\lambda_B, \lambda_M)$--plane for given $\mu$.} \end{figure} \vspace{0.8cm} \section{Conclusion} Our model has four parameters, $\,\lambda_M, \lambda_B, \mu, T\,$, but by scaling only their ratios are essential. For infinite temperature $\beta = 0\,$ eqs.(3.1--3) admit only the mean field solution $\Delta_B = 0\,,\, \Delta_M = \lambda_M,\,\, \overline W = \mu + \lambda_M$. This appears also from Fig. 5: since the phase structure of the model and the very existence of solutions depend on the ratios $\mu/T,\, \lambda_B/T,\, \lambda_M/T$, the area around the origin, so no superconducting phase present, reflects the situation for $T \rightarrow \infty$. By lowering the temperature one meets also the BCS--type solution but in a rather complicated region in the 3--dimensional parameter space. However, the solutions themselves in the limiting cases {\bf I.A}, {\bf II.A} are temperature independent while in {\bf I.B}, {\bf II.B} they do depend on $T$. Therefore, for given values of $\lambda_M, \lambda_B$ and $\mu$ the type--{\bf A} solutions do not change upon heating whereas the type--{\bf B} ones may appear and disappear, thus bringing the system through regions (a), (b) and (c) on Fig. 3(ii). Note that the limiting time evolution depends on the state. This would not be the case if norm convergence would be present as shown in \cite{NT}, but we have only strong resolvent convergence at our disposal. Whenever $\lambda_B$ is positive, it must be also $ \,>\mu$. Also for negative $\,\lambda_B,\, \lambda_M\,$ and $\,\lambda_M > -\mu\,$ there exists a finite gap for $\lambda_B$. A perturbation theory with respect to $\lambda_B$ is in general doomed to failure since for no point on the $\lambda_B = 0$ axis there is a neighbourhood full of the $\Delta_B \not= 0$ phase. It is interesting that without a mean field (the $\lambda_M = 0$ axis) there are superconducting solutions only for $\lambda_B > \mu$. An attractive mean field ($\lambda_M < 0$) stimulates superconductivity since then it also appears for negative $\lambda_B$. However, too strong mean field attraction destroyes it again. \section*{Acknowledgements} We are grateful to D.Ya. Petrina for stimulating our interest into the problem and to N.N. Bogoliubov Jr. for pointing out to us the absence of rigorous results to simultaneous mean--field -- BCS theory. We also appreciate suggestive discussions with H. Narnhofer. N.I. thanks the International Erwin Schr\"odinger Institute for Mathematical Physics where the research has been performed, for hospitality and financial support. This work has been supported in part also by ``Fonds zur F\"orderung der wissenschaftlichen Forschung in \"Osterreich" under grant P11287--PHY.
\section{Introduction} The standard sandpile model introduced in \cite{BTW} became of mathematical interest after the paper \cite{Dhar} by Dhar who discovered its Abelian structure. The model is defined on a finite hypercubic $d$-dimensional lattice. Each site $i$ is characterized by a nonnegative integer variable $z_i$ called the "height". If $z_i < 2d$ for all $i$, the sandpile is said to be stable. A vertex is picked at random and its height is increased by one. If $z_i \geq 2d$, then the site is unstable and topples giving one particle to each of its neighbors which in their turn, can be unstable and topple. The process called "avalanche" continues until all sites become stable and then a new particle is added to the lattice. If the avalanche reaches the boundary, sand disappears in a sink connected with all boundary sites. The "size" of an avalanche may be measured by the total number of topplings $s$, the number of distinct sites toppled $a$, the diameter of the region affected by avalanche $r$, and duration of avalanche $t$. It is generally believed, the probability that the avalanche has size $x (x=s,a,r,t)$ varies asymptotically as $x^{-\tau}(\tau=\tau_s,\tau_a,\tau_r,\tau_t)$ when $x \rightarrow \infty$. The mean-field value of the exponent $\tau_s$ obtained from exact solutions on the Bethe lattice \cite{DM} and on the full graph \cite{JL} is $3/2$. It is expected that $\tau_s=3/2$ also for ${\bf Z}^d$ when $d>d_u$ where $d_u$ is the upper critical dimension. The first attempt to find $d_u$ was made by Obukhov \cite{O} soon after the sandpile model was proposed. Briefly, Obukhov's arguments can be summarized as follows. (i) Consider changes resulting from an avalanche propagating through the lattice. If the system is in the recurrent state, the average sensitivity to a new excitation does not change. During the avalanche, the sites which have already toppled have heights lower than average, whereas the heights in the neighboring sites are larger than average. Hence, the previously activated sites repulse the new activation process. This situation resembles the True Self-Avoiding Walk (TSAW). (ii) The essential difference between the sandpile and TSAW is a possibility of branching the activation process. Thus, the avalanche can be pictured as a self-avoiding branching process. Using the renormalization group and $\epsilon$-expansion, Obukhov \cite{O} calculated one-loop corrections to the mean-field theory and came to the value $d_u=4$. This conclusion was supported by Dias-Guilera \cite{DG} who analyzed non-linear stochastic differential equations derived from the models with continuously distributed heights \cite{Zh}. Later on, Christensen and Olami \cite{CO} suggested $d_u=6$ from an analogy between spreading of avalanches and percolation. The mean-field treatment of a self-organized branching process was discussed in \cite{ZLS}. The proof of correspondence between the branching TSAW and an avalanche process in the sandpile model meets two considerable difficulties. First, a typical avalanche contains multiple topplings which violate self-avoidance of branches. In higher dimensions, multiple topplings are suppressed \cite{L},\cite{LU}. However, then, the second problem arises. The branching process corresponding to an avalanche depends deterministically on a recurrent configuration where it was initiated. All recurrent states in the Abelian sandpile model have equal probabilities \cite{Dhar}. If self-avoiding branching processes having equal number of steps, have equal statistical weights, they can be considered as lattice trees also known as branched polymers. The upper critical dimension of branched polymers $d_u=8$ was predicted in \cite{LI} and then rigorously determined in \cite{BFG},\cite{TH},\cite{HS}. Thus, avalanches either do not correspond to the branching TSAW due to multiple topplings, or do not have the specific TSAW weights due to equal probabilities of recurrent states. The computation situation is not less controversial. For the most part of numerical experiments, the accuracy is not sufficient to distinguish between non-trivial and mean-field values of critical exponents in high dimensions. Only high-statistics data on large lattices provide some information. Grassberger and Manna \cite{MG} investigated critical exponents for $d \leq 5$ and concluded that $d_u=4$. L{\"u}beck and Usadel \cite{LU} found $d_u=4$, mentioning that the largest considered size of the system $(L=80)$ for $d=4$ is too small and the corresponding avalanche distribution exhibits a very narrow power-law interval. In his paper \cite{L}, L{\"u}beck pointed an important role of logarithmic corrections to the scaling at $d_u=4$. At the same time, Chessa et al. \cite{CMVZ} performed a numerical study of critical exponents in dimensionality ranging from $d=2$ to $d=6$ and observed the mean-field behavior only in $d=6$ excluding, therefore, $d_u=4$. Very recently, Vespignani et al \cite{alex} have derived $d_u = 4$ from a phenomenological field theory, reflecting the symmetries and conservation laws of sandpiles. An apparent inconsistency between the simple self-avoiding branching process and real avalanches, as well as contradictory numerical results of different groups, set to find a more transparent proof of the upper critical dimension for sandpiles. In this paper, we prove the upper critical dimension $d_u=4$ using Lawler's theorems \cite{Law} for intersection probabilities of random walks and loop-erased random walks. We introduce again the self-avoiding branched polymers for description of avalanches and show that avalanches are spanning subtrees embedded into a spanning tree of the whole lattice rather than usual lattice trees. The problem of fractal dimension of avalanches is reformulated as that for the spanning subtrees. Using Majumdar's result \cite{M} of the equivalence between the chemical path on a spanning tree and the loop-erased random walk \cite{Law1}, we reduce the problem to estimations of intersection probability between random walks and loop-erased random walks. \section{The Model} We consider the sandpile model on the $d$-dimensional hypercube $\Lambda \subset {\bf Z}^d$. Elements of the state space $\{0,1,2,...(2d-1) \}^{\Lambda}$ are called stable configurations and are denoted by $C$. The value $C(i)=z_i, i \in \Lambda$ is the height of the sandpile at the site $i$. Given a configuration $C$ and a lattice site $i$, $a_i C$ is the stable configuration obtained by adding a particle at $i$, and relaxing the system by topplings at all unstable sites $j$, $z_j \geq 2d$. On toppling at the site $j$, \begin{equation} \label{1} z_{i} \rightarrow z_{i} - \Delta_{ij}\hspace{1cm}i \in \Lambda \end{equation} Elements of the matrix $\Delta$ are: $\Delta_{ii}=2d$ for all $i \in \Lambda$; $\Delta_{ij}=-1$ for all bonds $(i,j)$, $|i-j|=1$. The operators $a_i$ commute with each other \cite{Dhar} \begin{equation} \label{2} [a_{i},a_{j}]=0 \hspace{1cm}i,j \in \Lambda \end{equation} This allows one to define the identity operator \cite{Dhar} \begin{equation} \label{3} I_i = \prod_{j \in \Lambda} a_j^{\Delta_{ij}} \hspace{1cm} i \in \Lambda \end{equation} and the equivalence relation between two configurations $C^{'}$ and $C^{''}$ \begin{equation} \label{4} C^{'}(j) = C^{''}(j) + \sum_i n_i\Delta_{ij} \hspace{1cm} j \in \Lambda \end{equation} where $n_i, i \in \Lambda$ are integers. The equivalence means that both $C^{'}$ and $C^{''}$ tend to the same stable configuration after topplings at all unstable sites. The set of stable configurations $\{ C \}$ splits into two subsets: $\{ C \} = \{ C \}_R \bigcup \{ C \}_T$ where $\{ C \}_R$ is the recurrent set and $\{ C \}_T$ is the transient set of the sandpile process. In $\{ C \}_R$, one can define an inverse operator $a_i^{-1}, i \in \Lambda$. Then, the operators $\{ a_i \}$ form a finite Abelian group \cite{Dhar}. Due to Eq.(\ref{4}), each $C \in \{ C \}_R$ is an element of the super-lattice in the $|\Lambda|$-dimensional Euclidean space. The basis-vectors of this lattice are the rows of the matrix $\Delta$. Therefore, the number of elements in $\{ C \}_R$ is \cite{Dhar} \begin{equation} \label{5} N_R = Det \Delta \end{equation} The recurrent configurations in $\{ C \}_R$ have equal probabilities $N_R^{-1}$. Denote by $\partial \Lambda$ the set of boundary sites of $\Lambda$ and by $B$ the set of elementary cubes of the $d$-dimensional Euclidian space centered at $i \in \partial \Lambda$. We choose a site $\hat{i} \in {\bf Z}^d$ not belonging to $\Lambda$ and call it the root. The faces of cubes from $B$ which can be connected with $\hat{i}$ without intersections with another face, are called external faces. We connect each site $i \in \partial \Lambda$ with $\hat{i}$ by $\nu_i$ bonds where $\nu_i$ is the number of external faces of the cube $i$. The lattice $\Lambda$ together with $\hat{i}$ and new bonds forms a graph denoted by $G$. The site $\hat{i}$ is a sink for particles leaving the lattice $\Lambda$ during the avalanche. Using the explicit form of the identity operator Eq.(\ref{3}), one can construct a new identity operator \begin{equation} \label{6} I_{\partial \Lambda} = \prod_{i \in \partial \Lambda}a_i^{\nu_i} \end{equation} $I_{\partial \Lambda}$ says that adding $\nu_i$ particles to each boundary site triggers an avalanche of a special form: each site $i \in \Lambda$ topples exactly once and the initial configuration $C$ remains unchanged. One can use $I_{\partial \Lambda}$ to construct a graph representation for any $C \in \{ C \}_R$ by the so-called "burning algorithm" \cite{MD}. Consider the topplings initiated by $I_{\partial \Lambda}$ as a fire starting at $\hat{i}$ and burning sequentially all lattice sites. Once the rules of propagation of fire are fixed, the set of bonds along which fire propagates forms a spanning tree $T$ of the graph $G$. There is one-to-one correspondence between $ \{ C \}_R$ and the set of spanning trees $T$. Then, Dhar's formula, Eq.(\ref{5}), for $N_R$ coincides with Kirhhoff's theorem for the number of spanning trees \cite{Harary}. The matrix $\Delta$ in Eq.(\ref{5}) is the minor of the Laplacian matrix $\Delta_G$ of the graph $G$ corresponding to the element $(\Delta_G)_{\hat{i} \hat{i}}$. According to Dhar \cite{Dhar}, the Green function \begin{equation} \label{7} G_{ij} = [\Delta^{-1}]_{ij} \end{equation} is the average number of topplings at the site $j$ due to a single particle added at $i$ in a configuration $C \in \{ C \}_R$. For the random walk defined by the matrix $\Delta$, $G_{ij}$ is the expected number of times the walk started at $i$ visits the site $j$ until it is trapped by the absorbing site $\hat{i}$ \cite{Spitzer}. To find the spanning tree representation for $G_{ij}$, we delete from $\Lambda$ the elementary cube centered at $i$ and consider its $2d$ faces as a part of the new external boundary. The sites adjacent to $i$ form a new boundary set $\partial^i \Lambda$. Repeating the construction of a spanning tree by the identity operator, Eq.(\ref{6}), with the boundary set $\partial \Lambda \cup \partial^i \Lambda $, we obtain a two-component spanning tree having the roots $\hat{i}$ and $i$ for two different subtrees. {\it Proposition 1} proved in \cite{IKP} (see also \cite{IKPP}) reads: For any connected graph $G$ with a fixed vertex $\hat{i}$ \begin{equation} \label{8} G_{ij} = N^{(i,j)}/|T| \end{equation} where $N^{(i,j)}$ is the number of two-component spanning trees having the roots $\hat{i}$ and $i$, such that both vertices $i$ and $j$ belong to the same component; $|T|$ is the number of spanning trees on $G$. \section{Avalanches and Waves} An avalanche starting at $i$ is the process of transformation $C \rightarrow a_i C$. Generally, an avalanche consists of multiple topplings at different sites and has a complicated structure. However, one can try to decompose it into simpler subprocesses. Suppose $z_i = 2d$ after adding one particle to $i$. Since the topplings can be performed in any order, we topple once at $i$, and then topple all other unstable sites keeping the site $i$ out of the next toppling. This sub-avalanche is the first wave of topplings. If $i$ is still unstable, we topple $i$ once again giving rise to the second wave of topplings. This process is continued until $i$ becomes stable. Thus, an avalanche is broken into a sequence of waves of topplings. Three important properties of waves make them useful for the analysis of avalanche statistics \cite{IKP}. Let $S_k, k \geq 1$ be a set of sites toppled during the $k$-th wave. Then, (i) each $j \in S_k$ topples once and only once; (ii) both $S_k$ and $\Lambda \setminus S_k$ are connected sets; (iii) $k$-th wave is the last wave of an avalanche iff the initial site $i \in S_k$ has a neighbor in $\Lambda \setminus S_k$. The number of waves in the avalanche $C \rightarrow a_i C$ will be denoted by $n_i(C)$ or simply $n_i$. The spanning tree representation of waves can be constructed as follows. Given a configuration $C \in \{ C \}_R$, $P_i(C)$ represents the projection of $C$ on the lattice $\Lambda \setminus i$ where the site $i$ is now considered as the second root. Let $C_k, k \leq n_i$ be an (unstable) configuration obtained after the $k$-th wave starting at $i$ is completed. Then, $P_i(C_k)$ can be produced from $P_i(C)$ by the operator \begin{equation} \label{9} W_k = (\prod_{j\in\partial^i\Lambda}a_j)^k \end{equation} acting on the lattice $\Lambda \setminus i$. Consider the configuration \begin{equation} \label{10} P_i(C_{k-1}) = W_{k-1}P_i(C) \end{equation} preceeding the $k$-th wave. If one applies to $P_i(C_{k-1})$ the identity operator \begin{equation} \label{11} I_{\partial \Lambda \cup \partial^i \Lambda } = (\prod_{j\in\partial^i\Lambda}a_j) (\prod_{j\in\partial\Lambda}a_j^{\nu_j}), \end{equation} one obtains a two-component spanning tree representing $P_i(C_{k-1})$. On the other hand, the expression in the first brackets is exactly the operator providing the $k$-th wave. Therefore, the subtree rooted at $i$ and embedded into the subtree rooted at $\hat{i}$, is the graph representation of the $k$-th wave. This result gives another proof of Eq.(\ref{8}). As each wave beginning at $i$ and involving $j$ corresponds to one toppling at $j$, $N^{(i,j)}/|T|$ is the average number of topplings at the site $j$ for a single particle added at $i$. Then, Eq.(\ref{8}) follows from Eq.(\ref{7}). Due to the graph representation, we come again to a lattice-tree or a branched-polymer picture of avalanches. However, the essential difference with the qualitative Obuhkov arguments is, firstly, that we associate the lattice trees with waves but not with the whole avalanches, and secondly, the lattice trees are conditioned by the spanning tree construction. Let $s_k$ denote the number of sites in the $k$-th wave $S_k$. The total number of topplings $s$ in an avalanche is \begin{equation} \label{12} s = \sum_{k=1}^{n_i}s_k \end{equation} where $n$ is the number of topplings at the initial point $i$ or, equivalently, the number of waves in the avalanche. It is expected that, at least for sufficiently large dimensions $d$, the probability distribution of $s$ has the scaling form \begin{equation} \label{13} P(s;L) \sim s^{-\tau} f(s/L^D) \end{equation} where $D$ is the capacity fractal dimension of avalanches. It follows from Eq.(\ref{12}) that $D$ depends on the fractal dimension of waves and the number of waves in an avalanche. In $d=2$, the dimension of waves is $d_w = 2$ and the number of waves is expected to obey the scaling law \cite{MD} \begin{equation} \label{14} \langle n_i \rangle \sim r^y \end{equation} Then the exponents $\tau_s,\tau_a,\tau_r,\tau_t$ can be determined in terms of $y$. In higher dimensions, multiple toppling events occurs more rarely showing a tendency to decrease $y$. In Section {\bf 7}, we will show that for $d=4$, $\langle n_i \rangle$ grows with $r$ not faster than logarithmically. If so, the problem of critical dimension for avalanches can be formulated as that for waves. \section{Statistics of Waves} We start this section with the definition of the fractal dimension of waves. Consider the graph representation of a wave as a two-component spanning tree on the graph $G$ with the roots $i \in \Lambda$ and $\hat{i}$. The wave $S$ is the set of $s = |S|$ sites belonging to the component rooted at $i$. The radius of the wave is \begin{equation} \label{15} R(S) = sup\{ d(i,j):j \in S\} \end{equation} where $d(i,j)$ is the distance between the sites $i$ and $j$. We will say that waves have a fractal growth if \begin{equation} \label{16} \langle s \rangle \sim R^{d_{fg}} \end{equation} for large $R$, where $d_{fg}$ is the dimension of fractal growth and the average is taken over all waves of the radius $R$. The fracral growth, however, does not guarantee the fractal structure by itself. Surround the point $i$ by a ball of radius $r$ and consider all sites of $S$ inside the ball. Denote the number of internal points by $m(r)$. We say that waves have the fractal density if, given $\epsilon > 0$, there exist $r(\epsilon)$, $R(\epsilon)$ and a constant $\bar{c} < 1$ such that \begin{equation} \label{17} |log\langle m(r) \rangle -d_{fd} log r| \leq \epsilon \end{equation} on the interval $r(\epsilon) \leq r \leq \bar{c}R$ for all $R > R(\epsilon)$. If $d_{fg} = d_{fd} =d_f$, the fractal is uniform and $d_f$ is the fractal dimension of waves on the $d$-dimensional lattice. We do not prove here uniformity of the fractal structure of waves. Assuming it, we will find the upper and lower bounds for $d_f$. In the spherically symmetrical case, we can introduce the density of waves which varies as \begin{equation} \label{18} \rho(r) \sim \frac{1}{r^{d-d_f}} \end{equation} in the scaling interval $r(\epsilon) \leq r \leq \bar{c}R$. If $d_f$ is known, one can determine the probability distribution of waves $P(x), x=s,a,R,t $ ($s=a$ for waves). Indeed, the expected number of topplings at a point $j$ in an avalanche starting at $i$, $|i-j|=r$ varies as \begin{equation} \label{19} \langle n(r) \rangle \sim \rho(r) \int_{r/\bar{c}}^{\infty}P(x)dx \end{equation} where the integral gives all waves having reached the radius $r/\bar{c}$. On the other hand, $\langle n(r) \rangle$ coincides with the Green function, Eq.(\ref{7}). The asymptotics of $G(r)$ in the $d$-dimensional case is \begin{equation} \label{20} G(r) \sim \frac{1}{r^{d-2}} \end{equation} Comparing with Eq.(\ref{19}),we find that \begin{equation} \label{21} P(R) \sim \frac{1}{R^{d_f-1}} \end{equation} The probability distribution for $s$ follows from Eq.(\ref{16}) provided that $s \sim R^{d_f}$ when $R \rightarrow \infty$ \begin{equation} \label{22} P(s) \sim \frac{1}{s^{\tau_s}} \end{equation} with $\tau_s = 2-2/d_f$. The exponent $\tau_s$ reaches its mean-field value $3/2$ when $d_f = 4$. To proceed with the determination of $d_f$, we need a more elaborate definition of $\rho(r)$. Consider a wave $S$ with the initial site $i$. The set of sites not belonging to $S$ is denoted by $\hat{S}$, $S \cup \hat{S} = \Lambda$. According to Eq.(\ref{15}), a wave has the radius not less than $R$ if there exists a site $i^{'} \in S$, $|i-i^{'}| = R$. The wave $S$ has the density less than 1 at the radius $r \leq R$ if there exists a site $\hat{i}^{'} \in \hat{S}$, $|i-\hat{i}^{'}| = r$. Consider the two-component spanning tree with the components corresponding to $S$ and $\hat{S}$. The path $\Gamma(i,j)$ on a tree between the points $i$ and $j$ is the sequence of bonds $(i,i_1),(i_1,i_2),...(i_n,j)$. Any two points belonging to one component can be connected by a unique path. Then, we can define the density of waves at the radius $r \leq R$ by the conditional probability that, given a path $\Gamma(i,i^{'})$, $|i-i^{'}| = R$, there exists a path $\Gamma(\hat{i},\hat{i}^{'})$, $|i-\hat{i}^{'}| = r$, \begin{equation} \label{23} \rho(r) = 1 - Prob(\Gamma(\hat{i},\hat{i}^{'})|\Gamma(i,i^{'})) \end{equation} The expression in the right-hand side of Eq.(\ref{23}) is the non-intersection probability of two self-avoiding paths on a tree, called often "chemical paths". The first path $\Gamma(i,i^{'})$ connects the site $i$ with the site $i^{'}$ belonging to the wave $S$; the second path $\Gamma(\hat{i},\hat{i}^{'})$ connects the site $\hat{i}^{'}$ not belonging to $S$ with the sink $\hat{i}$. Having no rigorous results for this probability, we can, however, reduce it to non-intersection probability between one chemical path and the simple random walk. For this purpose, we need a generalization of Proposition 1 to the multicomponent case. Consider a connected graph $G$ and select a subset of its vertices $\{v\}=i_1,i_2,...,i_{\nu} $. Let $\Delta_{\{v\}}$ be the matrix obtained from the Laplacian matrix $\Delta_G$ by deleting all rows and columns corresponding to $i_1,i_2,...,i_{\nu} $. The Green function \begin{equation} \label{24} G_{ij}=[\Delta_{\{v\}}^{-1}]_{ij} \end{equation} is the expected number of visits of the site $j$ by the random walk starting at $i$ before it is absorbed at one of the sites $i_1,i_2,...,i_{\nu} $. {\it Proposition 2}\hspace{0.5cm} For any connected graph $G$ with a fixed subset of vertices $\{v\}$ \begin{equation} \label{25} G_{ij}=\frac{ N_{\{v\}}^{(ij)}}{|T_{\{v\}}|} \end{equation} where $ N_{\{v\}}^{(ij)}$ is the number of $(\nu+1)$-component spanning trees having the roots $i_1,i_2,...,i_{\nu}$ and $i$ such that vertices $i$ and $j$ belong to the same component; $|T_{\{v\}}|$ is the number of $\nu$-component spanning trees on $G$ having the roots $i_1,i_2,...,i_{\nu} $. {\it Remark} If a site $i_{\mu} \in \{v\}$ is an isolated site of the spanning tree, we consider it as a component consisting of the single site. {\it Proof} Consider the graph $G^{'}$ which is the graph $G$ with all sites $i_1,i_2,...,i_{\nu} $ contracted to the single site $i_0$. Proposition 1 is applicable to $G^{'}$. It relates the Green function, Eq.(\ref{24}), with the number of two-component spanning trees having the roots $i_0$ and $i$, such that the sites $i$ and $j$ belong to the same component. Returning to the graph $G$, we get from each configuration on $G^{'}$, a spanning tree where the component containing $i_0$ splits into $\nu$ components. This construction implies Proposition 2. In order to treat the right-hand side of Eq.(\ref{23}) as a random walk probability, we consider the graph $G$ with the subset of vertices $\{v\}$ coinciding with the set of all sites of the chemical path $\Gamma(ii^{'})$. Then, by Proposition 2, we have \begin{equation} \label{26} G_{\hat{i}^{'}\hat{i}}=\frac{ N_{\{v\}}^{(\hat{i}\hat{i}^{'})}}{|T_{\{v\}}|} \end{equation} where $G_{\hat{i}^{'}\hat{i}}$ is the expected number of visits of the site $\hat{i}$ by the random walk starting at $\hat{i}^{'}$ and escaping the chemical path $\Gamma(ii^{'})$. If $\nu$ is the number of sites in $\Gamma(ii^{'})$, then $N_{\{v\}}^{(\hat{i}\hat{i}^{'})}$ is the number of $(\nu+1)$ component spanning trees having sites $\hat{i}$ and $\hat{i}^{'}$ in one component. The remaining $\nu$ components can be joint into a single component if one adds to $\nu$ spanning subtrees rooted at the sites of $\Gamma(ii^{'})$, the path $\Gamma(ii^{'})$ itself. Then, $N_{\{v\}}^{(\hat{i}\hat{i}^{'})}$ is the number of two-component spanning trees which have the path $\Gamma(ii^{'})$ in one component corresponding to the wave $S$ and the path $\Gamma(\hat{i}\hat{i}^{'})$ in the second component $\hat{S}$ having the root $\hat{i}^{'}$ or, equivalently, $\hat{i}$. $|T_{\{v\}}|$ is the number of all one-component spanning trees containing the path $\Gamma(ii^{'})$. Along with $|T_{\{v\}}|$, we introduce $N_{\{v\}}^{(\hat{i})}$, the number of two-component spanning trees containing the path $\Gamma(ii^{'})$ in one component and the root $\hat{i}$ in the second one. By definition, \begin{equation} \label{27} Prob(\Gamma(\hat{i}\hat{i}^{'})|\Gamma(ii^{'}))= \frac{ N_{\{v\}}^{(\hat{i}\hat{i}^{'})}}{N_{\{v\}}^{(\hat{i})} }= \frac{N_{\{v\}}^{(\hat{i}\hat{i}^{'})}}{|T_{\{v\}}|} \frac{|T_{\{v\}}|}{N_{\{v\}}^{(\hat{i})}} \end{equation} Using Eq.(\ref{26}), we get \begin{equation} \label{28} Prob(\Gamma(\hat{i}\hat{i}^{'})|\Gamma(ii^{'}))= \frac{G_{\hat{i}^{'}\hat{i}}}{G_{\hat{i}\hat{i}}} \end{equation} where the Green function $G_{\hat{i}\hat{i}}$ is the expected number of returns to the initial point for random walks starting at $\hat{i}$ and escaping the path $\Gamma(ii^{'})$. Finally, we note that the ratio of Green functions in Eq.(\ref{28}) is the probability that the random walk starting at $\hat{i}^{'}$ escapes the path $\Gamma(ii^{'})$ and reaches the site $\hat{i}$. This fact follows from the property of Green functions known for infinite lattices as the "weak ergodic theorem" \cite{Spitzer}. Denoting by $F(\hat{i}^{'}|ii^{'})$ the escaping probability averaged over all paths $\Gamma(i,i^{'})$, we can summarize the obtained results as {\it Proposition 3}\hspace{0.5cm} The waves starting at a site $i$ of the lattice $\Lambda$ with the sink $\hat{i}$ have the density \begin{equation} \label{29} \rho(r) = 1 - F(\hat{i}^{'}|ii^{'}) \end{equation} at the point $\hat{i}^{'}$, $|i-\hat{i}^{'}|=r$, if the radius of the wave exceeds $R = |i-i^{'}| \geq r$. \section{Loop erased random walks} The loop erased random walk (LERW) has been introduced by Lawler \cite{Law1} to modify the uniform measure of the usual self-avoiding random walk where every possible walk of a fixed length is given the same statistical weight. Let $[x_{0},...,x_{m}]$ be the simple random walk in ${\bf Z}^{d}$. Then the LERW can be constructed as follows: let $j$ be the smallest value such that $x_{j}=x_{i}$ for $0\leq i < j \leq m$. Deleting all steps between $i$ and $j$, we obtain a new walk $[x_{0},...,x_{i},x_{i+1},...,x_{m}]$. If the new walk is self-avoiding we stop the process; otherwise, we continue the loop-erasing operation until we get a self-avoiding path. The distribution for each step of the LERW depends on the entire past history being equivalent to that given by the transition probability of the Laplacian random walk \cite{LE},\cite{Law2}. Another equivalence has been found by Majumdar \cite{M} who considered a model of growing trees introduced by Broder \cite{Brod}. The chemical path on these trees corresponds to the LERW. Like the burning algorithm, Broder's algorithm in the limit of lattice filling gives spanning trees with equal probability. Therefore, the statistical properties of the LERW are identical to those of the chemical path on spanning trees. Let $R_n(i)$ and $L_n(i)$ be simple and loop-erased random walks starting at the site $i$ and let $\Pi_i$ and $\Gamma_i$ denote paths of the walks \begin{equation} \label{30} \Pi_i(a)=\{R_n(i): 0\leq n \leq a\} \end{equation} \begin{equation} \label{31} \Gamma_i(a)=\{L_n(i): 0\leq n \leq a\} \end{equation} Denote by $\sigma(n)$ the number of steps of the simple random walk needed to produce $n$ steps of the LERW \begin{equation} \label{32} \sigma(n) = sup\{j:L_n(i)=R_j(i)\} \end{equation} For any random walk, $\sigma(n) \geq n$. For a fixed $\sigma(n)$, $\Pi_i(\sigma(n))$ is the path of a finite random walk. Then, $\Gamma_i(n)$ is the finite LERW obtained from $\Pi_i(\sigma(n))$ by the loop-erasing procedure. The escaping probability of a set $A \in {\bf Z}^d$ is defined as \begin{equation} \label{33} Es(i,A) = Prob \{R_n(i) \notin A,n=1,2,3,... \} \end{equation} where $i$ is the starting point of the random walk. Putting $A = \Gamma_i(n)$, we can define $a_n$ as \begin{equation} \label{34} a_n = E[Es(i,\Gamma_i(n))] \end{equation} i.e. $a_n$ is the probability that an infinite independent simple random walk starting in $i \in {\bf Z}^d$ does not intersect the finite LERW derived by erasing loops from $\sigma(n)$ steps of an independent random walk also starting in $i$. For $d=4$, Lawler \cite{Law53} has found the lower and upper bounds for $a_n$ \begin{equation} \label{35} \lim_{n\rightarrow \infty} inf\frac{\log a_n}{\log\log n}\geq -\frac{1}{2} \end{equation} and \begin{equation} \label{36} \lim_{n\rightarrow \infty} sup\frac{\log a_n}{\log\log n}\leq -\frac{1}{3} \end{equation} The upper bound Eq.(\ref{36}) was conjectured \cite{Law53} to coincide with the exact asymptotics. This result would imply that the mean square distance for LERW's behaves as \begin{equation} \label{37} \langle r_n^2\rangle \sim n (\log n)^{1/3} \end{equation} whereas $\langle r_n^2\rangle \sim n$ for $d \geq 5$. It is convenient to bring here the known intersection probability for two simple random walks \cite{Law}. Let $R_n(i)$ and $R_n(j)$ be independent random walks starting at the points $i$ and $j$, $|i-j|=r$. Then, there exist constants $c_1 > 0$ and $c_2 > 0$ such that \begin{equation} \label{38} c_1 \leq Prob\{\Pi_i(n) \cap \Pi_j(n)\neq 0\} \leq c_2 \end{equation} for $d<4$ \begin{equation} \label{39} c_1(\log n)^{-1} \leq Prob\{\Pi_i(n) \cap \Pi_j(n)\neq 0\} \leq c_2(\log n)^{-1} \end{equation} for $d=4$ \begin{equation} \label{40} c_1 n^{(4-d)/2} \leq Prob\{\Pi_i(n) \cap \Pi_j(n)\neq 0\} \leq c_2 n^{(4-d)/2} \end{equation} for $d>4$, if $a\sqrt n \leq |r| \leq b\sqrt n $, where $a$ and $b$ are positive constants. The non-intersection probability of a finite random walk and an infinite random walk starting at the same point $i$ for $d=4$ is \cite{Law},\cite{Dup} \begin{equation} \label{41} Prob\{\Pi_i(n) \cap \Pi_i(\infty) = 0\} \sim c(\log n)^{-1/2} \end{equation} Also, for further estimations, we need the intersection probabilities between a finite random walk starting at $i$ and an infinite random walk starting at $j$ for $d=4$ \cite{Law86} \begin{equation} \label{42} Prob\{\Pi_i(n) \cap \Pi_j(\infty) \neq 0\} \leq c_1 \log (1+1/\alpha)/\log n \end{equation} if $\alpha = |i-j|^2/n$. \section{Upper and lower bounds for density of waves} Every random walk on a graph $G$ with an absorbing set $\{v\}$ and the sink $\hat{i}$ is trapped either by $\{v\}$ or by $\hat{i}$. Then, Proposition 3 means that $\rho(r)$ is the intersection probability between the random walk starting at the site $\hat{i}^{'}, |i-\hat{i}^{'}|=r$ and the chemical path $\Gamma(i,i^{'}), |i-i^{'}|=R$. Majumdar \cite{M} has proved that chemical paths correspond to LERW's which, in turn, can be obtained from simple random walks by the loop-erasing procedure. Let $R_n(i)$ be the random walk corresponding to the LERW $L_n(i)$. As the number of steps $\sigma(n)$ in the path $\Pi_i(\sigma(n))$, Eq.(\ref{30}), needed to reach the radius $R$, exceeds the number of steps $n$ in the LERW path $\Gamma_i(n)$, Eq.(\ref{31}), we can estimate $\rho(r)$ as \begin{equation} \label{43} \rho(r) \leq Prob\{\Pi_i(\sigma(n)) \cap \Pi_{\hat{i}^{'}}(\infty)\neq 0\} \end{equation} In $d=4$, we get from Eq.(42) \begin{equation} \label{44} \rho(r) \leq c \log(1+\frac{1}{\alpha})/\log r \end{equation} where $\alpha = r^2/\sigma(n) \sim r^2/R^2$. The fractal density of waves in $d=4$ decays with $r$ at least logarithmically when $\alpha$ is fixed. In $d \geq 4$, we can use Eq.(\ref{40}) assuming that the asymptotics does not change when one of the random walks is extended to infinity. Then, we have \begin{equation} \label{45} \rho(r) \leq c r^{4-d} \end{equation} From the definition, Eq.(\ref{18}), we can see that the fractal dimension of waves $d_f \leq 4$ for all $d \geq 4$. In $d \leq 4$, the intersection probability, Eq.(\ref{38}), and, therefore $\rho(r)$, are restricted from above by a constant. To get a lower bound for $\rho(r)$ in $d=4$, we will use the upper bound Eq.(\ref{36}) for escaping probability $a_n$. Consider the random walk $R_n(i)$ and decompose its path $\Pi_i(\infty)$ into two parts: $\Pi_i(\infty) = \Pi_i(k) \cup \Pi_{\hat{i}^{'}}(\infty)$ where $k$ is the moment of the first hitting into the site $\hat{i}^{'}$ separated from $i$ by the distance $r$: $R_i(k) = \hat{i}^{'}, R_i(l) \neq \hat{i}^{'}, l=1,2,...,k-1 $. Then, escaping probability $a_n$ for the LERW of length $n$ by the random walk $R_n(i)$ is the product of escaping probabilities $a_n(i,\hat{i}^{'})$ and $a_n(\hat{i}^{'},\infty)$ before and after the first hitting into the site $\hat{i}^{'}$ \begin{equation} \label{46} a_n = a_n(i,\hat{i}^{'})a_n(\hat{i}^{'},\infty) \end{equation} Granting that $a_n(\hat{i}^{'},\infty) = F(\hat{i}^{'}|i i^{'})$, we can write Eq.(\ref{29}) as \begin{equation} \label{47} \rho(r) = 1 - \frac{a_n}{a_n(i,\hat{i}^{'})} \end{equation} where $n$ is the length of the LERW coinciding with the path $\Gamma(i,i^{'})$. The probability $a_n(i,\hat{i}^{'})$ to reach the point $\hat{i}^{'}$ from $i$ for $m$ steps, $m \sim |i-\hat{i}^{'}|^2=r^2$, escaping the LERW of length $n$ is not less than the probability to avoid the random walk of length $\sigma(n)$ from which the LERW was obtained by the loop- erasing procedure, and therefore not less than the probability $Prob\{ \Pi_i(\infty)\cap\Pi_i(m)=0\}$ to avoid an infinite random walk. Using Eqs.(\ref{36}) and (\ref{41}) we get from Eq.(\ref{47}) \begin{equation} \label{48} \rho(r) \geq 1 - c\frac{(\log m)^{1/2}}{(\log n)^{1/3}} \end{equation} For any $r \sim m^{1/2}$, we see from Eq.(\ref{37}) that the number of steps $n$ in the LERW increases with the radius of wave as $R^2$ (with the logarithmic correction) and $\rho(r)$ approaches 1 when $R \rightarrow \infty $. The only fractal dimension which is consistent with both upper and lower bounds Eq.(\ref{44}) and Eq.(\ref{48}) is 4 although the fractal density Eq.(\ref{18}) as well as the probability distribution of waves Eqs.(\ref{21}),(\ref{22}) need logarithmic corrections. The lower bound Eq.(\ref{48}) becomes stronger in $d < 4$ because of decreasing escaping probability $F(\hat{i}^{'}|ii^{'})$. Assuming that $d_f$ is a non-decreasing function of $d$, we conclude that the fractal dimension of waves is \begin{equation} \label{49} d_f= \left\{ \begin{minipage}{5cm} $d$ \hspace{1cm} for $d \leq 4$\\ 4 \hspace{1cm} for $d > 4$ \end{minipage} \right. \end{equation} According to Eq.(\ref{22}), the upper critical dimension for waves is $d_u = 4$. \section{Number of Waves in Avalanches} In order to derive the upper critical dimension for avalanches from that for waves, we need some information about the expected number of waves in an avalanche. Due to the property (iii) of waves mentioned in Section {\bf 3}, the probability that a wave starting at the site $i$ is the last wave in an avalanche is proportional to the density of "holes" in the wave at its origin, i.e. the probability that a site in the vicinity of $i$ does not belong to the wave. By Eq.(\ref{29}), this probability is $F(\hat{i}^{'}|ii^{'})$ which coincides simply with the escaping probability $a_n$ in the case $r=|i-\hat{i}^{'}|\rightarrow 0$. Avalanches of radius $r$ consist of the waves whose radius does not exceed $r$. Therefore, the expected number of waves can be estimated as \begin{equation} \label{50} \langle n_i \rangle \leq c a_n^{-1} \end{equation} where $r^2 \sim n(\log n)^{1/3}$. Using Eq.(\ref{35}), we get \begin{equation} \label{51} \langle n_i \rangle \leq c(\log n)^{1/2} \end{equation} or \begin{equation} \label{52} \langle n_i \rangle \leq c_1(\log r)^{1/2} \end{equation} Thus, in $d=4$, the number of waves in an avalanche of radius $r$ grows not faster than logarithmically. In $d=2$, the distribution of last waves is known \cite{DManna} \begin{equation} \label{53} P_l(R) \sim \frac{1}{R^{7/4}} \end{equation} The probability for the wave of radius $R$ to be the last in an avalanche is $P_l(R)$ divided by the general distribution of waves Eq.(\ref{21}). Then, the expected number of waves $\langle n_i \rangle$ in $d=2$ has the upper estimate \begin{equation} \label{54} \langle n_i \rangle \leq c r^{3/4} \end{equation} which is consistent with Eq.(\ref{14}) where the exact value $y=1/2$ was conjectured \cite{PKI}. For $d > 4$, the escaping probability $a_n$ remains finite for all $n$. Therefore, the expected number of waves in an avalanche is restricted by a constant. The logarithmic growth of $\langle n_i \rangle$ in $d=4$ given by Eq.(\ref{52}) implies that the probability distributions for waves and avalanches can differ not more than by logarithmic correction. To estimate this correction, we take instead of the density $\rho(r)$ its upper bound Eq.(\ref{44}) if r.f.s. Eq.(\ref{44})is less than 1, and put $\rho(r)=1$ otherwise. Comparing Eq.(\ref{20}) with Eq.(\ref{19}) where the new $\rho(r)$ is taken, we get the probability distribution for waves in the form \begin{equation} \label{55} P(R) = \frac{\log R}{R^3} \end{equation} Also, we take instead of the lower bound Eq.(\ref{35}) for $a_n$ its exact asymptotics conjectured by Lawler \cite{Law53}. Then, \begin{equation} \label{56} \langle n_i \rangle \sim (\log r)^{1/3} \end{equation} The maximal difference between avalanche and wave distributions corresponds to a situation when all waves in an avalanche of radius $r$ have the maximal radius $r$ \begin{equation} \label{57} \langle n_i \rangle P_{aval}(r)dr = P(r)dr \end{equation} Hence, the avalanche distribution in $d=4$ can be estimated as \begin{equation} \label{58} P_{aval}(r) \sim \frac{(\log r)^{\gamma}}{r^3} \end{equation} with $2/3 \leq \gamma \leq 1$. The probability distribution of the total number of topplings in a wave $P(s)$ follows from Eq.(\ref{55}) provided the leading asymptotic of $s$ is $s \sim R^4/\log R$: \begin{equation} \label{59} P(s) \sim \frac{(\log s)^{1/2}}{s^{3/2}} \end{equation} In $d > 4$, the asymptotics of the avalanche and wave distributions coincide and correspond to the mean-field behavior with the exponent $\tau_s = 3/2$. \section{Acknowledgements} This work was supported by RFFR through Grant No. 99-01-00882 and by NREL through Grant No. AAX-8-18685-01. The support and hospitality of the Dublin Institute of Advanced Studies are gratefully acknowledged . \newpage
\section{Introduction} In the last two decades in Quantum Optics there has been a fast and exciting development. Many experiments involving the optical field fall into the quantum domain, and many quantum mechanical {\em gedanken} experiments are now performed in optical laboratories. Basic concepts of quantum mechanics such as the reduction postulate, the uncertainty relation, the Sch\"odinger cat and nonlocal phenomena play a major role and their effects can be directly observed. Besides this fundamental interest, the field receives attention also in view of future applications, which are mainly motivated by the potential improvement offered by Quantum Mechanics to the manipulation and the transmission of information. Such recent developments have renewed the interest for two fundamental themes in Quantum Optics, namely the {\it generation} and the {\it measurement} of nonclassical states of the radiation field. \par In this paper, we address both these aspects and suggest a novel all-optical device which is able to select a fixed Fock component from a traveling wave initially prepared in a generic (possibly mixed) state. The scheme, which we named {\em Fock Filter}, consists of a ring cavity coupled to the signal through a cross-Kerr medium. At the output of the device, the signal and the cavity modes are strongly entangled, so that a successful photodetection of the cavity modes reduces the signal into a state with the desired number of photons. The proposed setup has two main applications. On one hand, it can be used to synthesize a generic number state $|n\rangle$ or a superposition of few number states, say $|\psi\rangle\propto\alpha |n_1 \rangle+ \beta |n_2\rangle $, starting from a coherent source. On the other hand, it allows to measure the photon distribution $P(n)=\langle n|\hat\nu |n\rangle$ and the density matrix $\hat\nu_{nk}=\langle n|\hat\nu |k\rangle$ of a generic input signal $\hat\nu$. \section{The Fock Filter} The device we propose is schematically depicted in Fig. 1. It consists of an active ring cavity coupled to the signal by a cross-Kerr medium. The cavity has a high quality factor, namely it is built by low transmissivity (denoted by $\tau$) beam splitters. The cross-Kerr interaction couples the cavity mode $d$ to the traveling signal mode $c_1$, according to the unitary evolution $\hat U_{K}=\exp(-i\chi t d^\dag d c_1^\dag c_1)$. In this way, the cavity mode experiences a phase-shift which depends on the quantum state of the signal mode. A further tunable phase-shift $\psi$ is also inserted in the cavity path. A port of the cavity (mode $a_1$) is fed by a strong coherent probe, whereas the second port (mode $a_2$) is left vacuum. At the output of the cavity, the mode $b_1$ is simply absorbed, whereas the mode $b_2$ is monitored by an avalanche photodetector. As we will see in the following, we only need to know whether or not any photon is present, namely to perform an {\sf ON/OFF} photodetection. \begin{figure}[h] \begin{minipage}{70mm} \psfig{file=stp.ps,width=64mm} \end{minipage}\hspace{8pt}\begin{minipage}{65mm} \caption{Fig.1: Schematic diagram of the Fock Filter. BS$_1$ and BS$_2$ denote low transmissivity beam splitters, and $\psi$ a tunable phase shift. The cavity input modes $a_1$ and $a_2$ are in a coherent state and in the vacuum respectively. The box denotes the cross Kerr medium that couples the cavity mode with the signal mode. A successful detection at the photodiode D filters the desired Fock component from the input signal. \label{f:stp}}\end{minipage} \end{figure}\par The probability operator measure (POM) describing such a detection scheme is given by the the two-value operator $\hat\Pi_0\doteq\sum_{k=0}^\infty(1-\eta)^k|k\rangle\langle k|$ and $\hat\Pi_1\doteq \widehat{{\Bbb I}} -\hat\Pi_0$ where $\eta$ is the quantum efficiency of the photodetector. The modes transformation of the cavity is expressed by [1] \begin{eqnarray} \left\{\begin{array}{l}b_1=\kappa(\varphi) a_1+e^{i\varphi}\sigma(\varphi) a_2\\ b_2=\sigma(\varphi) a_1+\kappa(\varphi) a_2\end{array}\right. \;,\label{ring} \end{eqnarray} where the phase-dependent transmissivity $\sigma$ and reflectivity $\kappa$ amplitude are given by \begin{eqnarray} \kappa(\varphi)\doteq\frac{\root\of{1-\tau}(e^{i\varphi}-1)} {1-[1-\tau]e^{i\varphi}}\qquad \sigma(\varphi)\doteq\frac\tau{1- [1-\tau]e^{i\varphi}},\label{defsigma} \end{eqnarray} with $\left|\kappa(\varphi)\right|^2+\left|\sigma(\varphi)\right|^2 =1$. The phase $\varphi$ is the total phase-shift experienced by the cavity mode, namely the sum of the shift due to the Kerr interaction and the tunable shift $\psi$. For the signal in a number state $|n\rangle$ the total phase shift imposed to the cavity mode is $\varphi\equiv \varphi_n =\psi - \chi n t$. In order to simplify notation, we write $\sigma _n\doteq \sigma(\varphi _n)$ and analogously $\kappa_n \doteq \kappa (\varphi_n)$. The input state of the device can be written as $ \hat\varrho_{in}=|\alpha\rangle\langle\alpha|\otimes|0\rangle\langle 0| \otimes\hat\nu_{in}$, namely a generic state $\hat\nu_{in}$ for the signal mode and a strong coherent state $|\alpha\rangle$ for the probe, the second port of the cavity being left unexcited. The output state can be easily found in the Schr\"odinger picture as \begin{eqnarray} \hat\varrho_{out} = \label{rhoout} \sum_{n,m=0}^{\infty} \nu_{nm} \: |\kappa_n\alpha\rangle \langle \kappa_m\alpha|\otimes|\sigma_n\alpha\rangle\langle \sigma_m\alpha|\otimes |n\rangle\langle m|\;. \end{eqnarray} The measurement scheme consists in detecting whether (detector D on) or not (detector D off) any photon is present at the output of the cavity. The corresponding probabilities are given by \begin{eqnarray} P_1={\mbox Tr}[\hat\Pi_1\hat\varrho_{out}] = \sum_{n=0}^\infty \nu_{nn} \left(1- e^{-\eta|\alpha|^2\:|\sigma_n|^2}\right) \qquad P_0 = 1-P_1 \label{P1} \;, \end{eqnarray} whereas the output signal conditioned by the photodetection reads \begin{eqnarray} \hat\nu_{out} ({\sf ON}) &=& \frac{e^{-|\alpha|^2}}{P_1} \sum_{n,m=0}^\infty \nu_{nm}\: e^{|\alpha|^2 \big[\kappa_n\kappa^*_m + \sigma_n\sigma^*_m \big]}\left(1- e^{-\eta|\alpha|^2 \sigma_n\sigma^*_m} \right)\:|n\rangle\langle m| \label{nuon} \end{eqnarray} The filtering properties of the device are due to the strong dependence of the cavity transmissivity on the internal phase-shift. The overall transmissivity function writes as \begin{eqnarray} |\sigma_n|^2 = \left[1+ 4 \frac{1-\tau}{\tau^2}\sin^2\frac{\psi-\chi nt}{2} \right]^{-1}\label{sigma}\;, \end{eqnarray} and exhibits a periodic structure sharply peaked at $n = n^*+ 2\pi j / (\chi t)$ with $n^*=\psi/(\chi t)$ and $j\in{\mathbb Z}$. In the peaks, it has unit height and width of the order of the beam splitter transmissivity $\tau$ (typically $\tau\sim 1\% - 0.01\%$). The value $n^*$ can be adjusted to an arbitrary integer by tuning the phase-shift $\psi$ as a multiple of $\chi t$, whereas multiple resonances are avoided by using relatively small values of the nonlinearity $\chi t$, so that the values of $n$ for $j\not = 0$ correspond to vanishing matrix elements $\nu_{ni}\simeq 0\ \forall i$. In this way the cavity is set at resonance only by a single Fock component $|n^*\rangle$ of the signal, which is {\em filtered} at the output in the case of successful photodetection. In the next section we will analyze this process in more detail. \section{Synthesis of number states} Let us now consider a cavity with a high quality factor (i.e. $\tau \ll 1$) adjusted to select the Fock component $n^*$ by tuning $\psi=\chi t n^*$. In this case, the detection probability and the conditional output states of Eqs. (\ref{P1}) and (\ref{nuon}) rewrite as follows \begin{eqnarray} P_1 \simeq \nu_{n^*n^*}+\frac{\eta |\alpha |^2 \tau^2}{(\chi t)^2}\sum_{p\neq n^*}\frac{\nu_{pp}}{(n^* -p)^2} \;,\label{P1asym} \end{eqnarray} and \begin{eqnarray} \hat\nu_{out} ({\sf ON}) \simeq \frac{1}{\sqrt{\cal N}}\left[ |n^*\rangle \langle n^*|+\frac{\eta |\alpha|^2 \tau^2}{(\chi t)^2}\sum_{n,k\neq n^*} \frac{\nu_{nk}}{(n^*-n)(n^*-k)}|n\rangle\langle k|\right]\;,\label{nuonasym} \end{eqnarray} where ${\cal N} = 1 + (\eta |\alpha|^2 \tau^2)/(\chi t)^2 \sum_{p\neq n^*} \nu_{pp}/(p-n^*)^2$ is a normalization constant. Both equations are valid when small values of the nonlinearity are involved, namely when a single Fock component sets the cavity into resonance. The physical meaning of Eqs. (\ref{P1asym}) and (\ref{nuonasym}) is apparent: when the cavity is "good enough" to appreciate the phase-shift imposed by the passage of the desired Fock component [i.e. when $\tau \ll (\chi t)$] the detection probability equals the probability of having $n^*$ photons in the signal $P_1 \simeq \nu_{n^*n^*}$, and the conditional output state approaches the corresponding number state $\hat\nu_{out} ({\sf ON}) \simeq |n^*\rangle\langle n^*|$, which is {\em synthesized} from the input signal. Eqs. (\ref{P1asym}) and (\ref{nuonasym}) also illustrate the effect of nonunit quantum efficiency of the probe photodetector. If $\eta$ is lower than $100\%$, the detection probability decreases, and thus also the synthesizing rate. However, the synthesized state is closer to the desired number state, namely the synthesizing quality is improved. In Fig. 2 the synthesis of the number state $|n^* \equiv 4\rangle$ is illustrated for decreasing values of the beam splitter transmissivity. \begin{figure}[h] \begin{tabular}{ccc} \psfig{file=num1.ps,width=4cm}&\psfig{file=num2.ps,width=4cm}& \psfig{file=num3.ps,width=4cm}\end{tabular} \caption{Fig.2: The photon distribution of the conditional output state $\hat \nu_{out} ({\sf ON})$ for different values of the transmissivity $\tau$, reported on each plot. The plots refer to the synthesis of the number state $|n^* \equiv 4 \rangle$, with $\chi t = 0.01$, $\psi=0.04$ and $\eta=0.8$. Both the probe mode $a_1$ and the signal mode $c_1$ are excited in a coherent state, with real amplitude $\alpha =20$ and $\beta =2$ respectively. The probabilities of obtaining the three states (i.e. the detection probability at D) are also indicated on each plot. The last plot shows that a transmissivity $\tau=0.02\%$ is sufficient for a reliable synthesis of the desired number state.} \end{figure} \par The Fock Filter behaves differently when larger nonlinearities or quite excited input signals are involved. In this case, the cavity may be set at resonance by several Fock components of the signal mode, corresponding to different integers that are multiple of $\chi t$. Let us consider, as an example, the situation in which two Fock components, corresponding to the values $n_1\equiv n^*$ and $n_2=n_1+2\pi/(\chi t)$, set the cavity into resonance. For $\tau \ll \chi t$ Eqs. (\ref{P1}) and (\ref{nuon}) can be written as $P_1 \simeq \nu_{n_1n_1}+ \nu_{n_2n_2}$ and $\hat\nu_{out}({\sf ON})\simeq 1/P_1 \Big[\nu_{n_1 n_1}|n_1\rangle\langle n_1|+\nu_{n_2 n_2}|n_2 \rangle\langle n_2| + \nu_{n_1 n_2}|n_1 \rangle\langle n_2|+\nu_{n_2 n_1}|n_2 \rangle\langle n_1|\Big]$. If the input signal $\hat\nu$ is excited in a coherent state, one has $\nu_{n_1 n_1}\nu_{n_2 n_2}=\nu_{n_1 n_2}\nu_{n_2 n_1}$, and hence $\hat\nu_{out}({\sf ON})$ is a pure state. Actually, by varying the amplitude of the input coherent signal any superposition of the form $|\psi\rangle \propto \alpha |n_1\rangle+ \beta |n_2\rangle$ may be synthesized at the output. We just mention that this kind of superposition is the paradigm for the realization of an optical qubit. \section{State measurement} In this section we show how the Fock filter can be used to measure the photon distribution, and the whole density matrix, of a generic input signal. The method is based on Eq. (\ref{P1}), which shows that for moderate nonlinearities and low beam splitter transmissivity ($\tau \ll \chi t < 1$) the detection probability $P_1$ at the output of the cavity is equal to the diagonal matrix element of the signal $P_1\simeq\nu_{n^*n^*}$ corresponding to the integer $n^*$ selected by tuning the phase $\psi$. Therefore, by repeated preparations of the signal and by varying the cavity tuning $\psi = \chi t n^*$, $n^*=0,1,2,...$ in order to span the whole excitation spectrum of the signal it is possible to record the photon number distribution of a generic input state. Actually, such a measurement may be implemented by a more efficient scheme using a set of cavities in cascade, each tuned on a different integer $n_k$. \begin{figure}[h] \begin{tabular}{ccc} \psfig{file=pdn1.ps,width=4cm}&\psfig{file=pdn2.ps,width=4cm}& \psfig{file=pdn3.ps,width=4cm}\end{tabular} \caption{Fig.3: Monte Carlo simulations of the detection of the photon number distribution by the Fock filter with $\chi t=0.1$, $\tau =0.1 \%$ and $\eta = 40 \%$. The distributions for a squeezed vacuum with $\langle a^\dag a\rangle =1$ average photons, a coherent state with $\langle a^\dag a\rangle=2$ average photons and a thermal state with $\langle a^\dag a\rangle=1$ average photons are reported from the left to the right together with the corresponding confidence interval. The empty squares indicates the theoretical values. In all cases a sample of $2000$ data has been used. } \end{figure} \par The input state of the $k$-th cavity is the output state from the $(k-1)$-th one and this allows to largely reduce the number of repeated preparations of the signal. At the $k$-th step, in the limit of good cavities, the detection probabilities approaches $P_1^{(k)} \simeq \nu_{n_kn_k}^{(k)}$ and $P_0^{(k)} \simeq 1- \nu_{n_kn_k}^{(k)}$ where the density matrix $\hat\nu^{(k)}$ has been reduced according to the result of the detection at the $(k-1)$-th photodiode as in Eq.(\ref{nuon}). For good cavities one has \begin{eqnarray} \hat\nu_{out}^{(k)}({\sf ON}) \simeq |n_k\rangle\langle n_k|\;, \qquad \hat\nu_{out}^{(k)}({\sf OFF})\simeq \sum_{p\neq n_k}\frac{\nu_{pp}^{(k)}} {P_0^{(k)}} |p\rangle\langle p| \label{nuth}\;. \end{eqnarray} We checked the whole detection strategy with a Monte Carlo simulation. In Fig. 3 we show the photon distributions obtained for a squeezed vacuum, a coherent state and a thermal state at the input. Remarkably, the photon distributions are reliably determined using a relatively small sample of data and a low value for the quantum efficiency of the photodetectors. \par The Fock filter, in conjunction with the unbalanced homodyning technique [2-3], allows also to measure the entire density matrix of the input signal. In order to achieve this goal, it is necessary to mix the input signal $\hat\nu$ with a strong coherent state $|z\rangle$ by using a high transmissivity beam splitter before the set of cavities (see Fig. 4). In the limit $(1-\tau) \ll 1$ and $|z| \gg 1$ the state entering the set of cavities is the displaced signal $\hat\nu_\gamma =\hat D(\gamma)\hat\nu\hat D^\dag (\gamma)$, where $\hat D (\gamma)=\exp(\gamma a^\dag - \hat\gamma a)$ is the displacement operator and $\gamma = z\sqrt{1-\tau}$. In this case, the Fock filter provides the photon distribution of the displaced state $P_\gamma (n)=\langle n|\hat \nu_\gamma|n\rangle$, which can be expressed in terms of the density matrix of the original signal $\hat\nu$ by the linear relation $ P_\gamma(n)= \sum_{km} \langle k|\hat\nu|m\rangle A_{kmn}(\gamma )$ where $A_{kmn}(\gamma )=\langle n|\hat D(\gamma)|k\rangle\langle m|\hat D^\dag (\gamma)| n\rangle$. By measuring the photon distribution for different values of the displacing amplitude $\gamma$, this relation can be inverted, leading to the reconstruction of the signal density matrix. \begin{figure}[h]\begin{minipage}{83mm} \psfig{file=cavdispl.ps,width=79mm}\end{minipage}\begin{minipage}{50mm} \caption{Fig.4: Schematic diagram of the setup for measuring the density matrix with the Fock filter. The input state is mixed with a strong coherent state by using a high transmissivity beam splitter, and then enters the set of cavities in cascade, each tuned on a different Fock component.}\end{minipage} \end{figure} \par In particular, Opatrny et al [4] have shown that it is enough to measure $P_\gamma (n)$ for a fixed value of the modulus $|\gamma |$ and different values of the phase. In this case one has \begin{eqnarray} P_{|\gamma|}^{(s)} (n) = \sum_{m=0}^{M-s} A_{m+s,m,n}(|\gamma |) \:\nu_{m+s,m} \label{tom1}\;, \end{eqnarray} where $M$ is the maximum Fock component excited in the signal, and $P_{|\gamma|}^{(s)} (n)$ is the Fourier transform of the photon distribution obtained by varying the phase $\varphi =\arg (\gamma$) The linear system (\ref{tom1}) is overdetermined and may be solved by least squares method. The solution represents the best estimate for the density matrix of the input signal. \section{Conclusions} In this paper, we have studied a novel all-optical device: the Fock filter, which is able to select the desired Fock component starting from a generic input state. The device may be used to synthesize number states and superpositions of few number states, as well as for measuring the photon distribution and the density matrix of a generic input signal. The feasibility of the proposed setup relies on the realization of cavities with a high quality factor, namely cavities built with low transmissivity beam splitters. Monte Carlo simulations have shown that transmissivities of the order of $\tau \sim 1\% - 0.01 \%$ are needed, which corresponds to beam splitters currently available in optical labs. We conclude by pointing out that the quantum efficiency of the photodetectors is not a crucial parameter for the Fock filter: low quantum efficiency does not degrade the performances of the device, for both the generation and the measurement scheme. \section*{Acknowledgments} This work is part of the INFM project PRA-CAT-97 and the MURST "Cofinanziamento 1997". MGAP thanks {\em Accademia Nazionale dei Lincei} for partial support through the {\em Giuseppe Borgia} award. \section*{References} \begin{description} \item{[1]} G. M. D'Ariano, L. Maccone, M. G. A. Paris,and M. F. Sacchi, unpublished \item{[2]} \refer{S. Wallentowitz, W. Vogel}{Phys. Rev. A}{53}{1996}{4528} \item{[3]} \refer{K. Banaszek, K. W\`odkievicz}{Phys. Rev. Lett}{76}{1996}{4344} \item{[4]} \refer{T. Opatrn\`y, D.-G. Welsch }{Phys. Rev. A} {55}{1997}{1462} \end{description} \end{document}
\section{Introduction} \paragraph In recent years there has been considerable activity in the study of random magnetic Schr\"odinger operators mainly due to their relation with the theory of the Integer Quantum Hall Effect (IQHE). Some of these studies have incorporated the randomness into the magnetic field$^1$, whereas others have added a random potential to the usual Landau Hamiltonian. Without any disorder the Landau Hamiltonian has a spectrum of evenly spaced {\sl Landau levels}, each one of which is an infinitely degenerate eigenenergy. When a random potential is added these Landau levels broaden into bands. In several models$^{2-4}$ it has been shown that for large magnetic field the spectrum at the edges of the bands is pure point, with each eigenenergy corresponding to an exponentially localized state. The proofs rely on von Dreifus and Klein's$^5$ refined version of the earlier multiscale analysis by Fr\"ohlich and Spencer$^6$ and on percolation theory. These results are not sufficient to provide a complete understanding of the IQHE however, as the nature of the spectrum in the interior of the band is crucial in explaining the observed plateaux$^7$. In one special case$^{8,9}$ the spectrum has been completely characterized. In this work the authors consider a random potential consisting of zero-range scatterers (delta functions) situated on the sites of a regular lattice. In the first paper$^8$, they show that, in the case of a single-band approximation, the whole spectrum is pure point, with exponentially localized states for all energies except the original Landau level. They prove also that this level remains infinitely degenerate. These results are improved in a later work$^9$ where they obtain similar results for the unprojected Hamiltonian. They adopt a simple proof of localization by Aizenman and Molchanov$^{10}$ which utilizes low moments of the resolvent kernel. \paragraph The purpose of this work is to generalize the above for the case of a magnetic Schr\"odinger operator with randomly distributed delta impurities. Specifically, the random potential consists of point scatterers, delta functions, positioned in unit squares which are centered on the Gaussian integers, so that it is possible to have up to four scatterers arbitrarily close together. The strengths of the scatterers will also be random. We consider a two-dimensional infinite system of noninteracting electrons moving in a uniform magnetic field of strength $B$ and the random potential $V$. The precise hypotheses on the probability distributions will be stated in Section II. In the symmetric gauge the vector potential is given by ${\bf A}({\bf r})={1\over 2}({\bf r}\times {\bf B})$ and the Hamiltonian is \begin{equation}\label{opb:fullham} H=(-i\nabla -{\bf A}({\bf r}))^2+V({\bf r}). \end{equation} When the magnetic field is sufficiently strong in comparison to the random potential the Landau bands do not overlap and it is sufficient to consider the projection of the Hamiltonian onto only one of them. The Hamiltonian retricted to the $n$th level is \begin{equation}\label{opb:projham} H_n=B(2n+1)P_n+P_nVP_n, \end{equation} where $P_n$ denotes the projection onto the level. The first term comes from the decomposition of the purely kinetic part of (\ref{opb:fullham}) and can be dropped as it modifies the energy only by a constant. Note that the resulting Hamiltonian is a random integral operator instead of a differential operator and that the kernels of $P_n$ are known explicitly. For simplicity, in this paper, we restrict ourselves to the case $n=0$ but the case $n\neq 0$ can be treated similarly. \paragraph For our model, in the special case where the support of the positional probability distribution is bounded within each unit box so that a corridor exists between impurities, the method of Aizenman and Molchanov yields a simple proof of localization$^{11}$. However, for general distributions of position, impurities can become arbitrarily close to each other and we are not able to use their method. This is partly due to possible resonances; that when impurities can become arbitrarily close together the low moments of the resolvent kernel do not converge rapidly enough. In this paper we use the modification of the Theorem of von Dreifus and Klein given in Ref. 2 to show exponential localization of states corresponding to each of the eigenenergies separately (except the original Landau level eigenenergy). We do this by studying, at fixed energy, the behaviour of the generalized eigenfunctions at the impurity sites only, thus reducing the problem to the study of a random matrix. The eigenfunctions of this matrix are related to the eigenfunctions of the Hamiltonian in such a way that exponential decay of the former implies exponential decay of the latter. Then using Kotani's `trick' $^{12}$ we can show exponential decay for all eigenenergies in an interval with probability one implying that the whole spectrum is pure-point. That the original Landau level eigenenergy remains infinitely degenerate has been shown in Ref. 13 for the case of a Poisson distribution of impurities. The result given here is similar and so only a sketch of the proof is given. The paper is organised as follows. In Section II we give a precise definition of the model. In Section III we characterize the spectrum as a set, state the main theorem and show infinite degeneracy of the original eigenenergy. Also in this section we relate the Hamiltonian to a lattice operator and state our version of the adapted von Dreifus-Klein Theorem. Section IV contains the main work of this paper, where the conditions of the main theorem are checked. In Section V we use Kotani's trick to show exponential decay and pure-point spectrum with probability one. \section{Definition and Boundedness of the Hamiltonian} \paragraph Let $\omega_n$, $n\in\hbox{\BB Z}[i]\equiv\{ n_1 +in_2 :\ (n_1,n_2)\in\hbox{\BB Z}^2\}$, the Gaussian integers, be independent, identically distributed (i.i.d) random variables representing the strengths of the impurities. We shall assume that their distribution is given by an absolutely continuous probability measure $\mu$ whose support is a compact interval $X=[a,b]\subset\hbox{\BB R}$ containing the origin and whose density $\rho $ is bounded by a constant $\rho_b$. We let $\Omega_1 =X^{\hbox{\BB Z}[i]}$ and $\hbox{\BB P}_1 =\prod_{n\in\hbox{\BB Z}[i]}\mu$. Define the unit squares centred at $n\in\hbox{\BB Z} [i]$: $$ B_n =\{ z\in\hbox{\BB R}^2\,\vert\, n_i-{1\over 2}\le z_i < n_i+{1\over 2}\,,\,n\in\hbox{\BB Z} [i]\,,\,i=1,2\,\}. $$ Let $\zeta_n =n+ {\tilde \zeta}_n$, $n\in\hbox{\BB Z}[i]$, represent the positions of the impurities in the complex plane. ${\tilde \zeta }_n$, $n\in\hbox{\BB Z} [i]$ are i.i.d. random variables. We shall assume that their distribution is given by a probability measure $\nu$ with support equal to $B_0$ and density $r $ bounded by a constant $r_b$. We let $\Omega_2 =\times_{n\in\hbox{\BB Z}[i]}B_0$ and $\hbox{\BB P}_2 =\prod_{n\in\hbox{\BB Z}[i]}\nu$. Our probability space will be $\Omega =\Omega_1\times\Omega_2$ with probability measure $\hbox{\BB P} =\hbox{\BB P}_1\times\hbox{\BB P}_2$. For $m\in\hbox{\BB Z} [i]$ let $\tau_m$ be the measure preserving automorphism of $\Omega$ corresponding to translation by $m$: \begin{equation} (\tau_m(\omega,\zeta))_n=(\omega_{n-m},\zeta_{n-m})\ . \end{equation} The group $\{\tau_m:m\in\hbox{\BB Z} [i]\}$ is ergodic for the probability measure $\hbox{\BB P}$. Let ${\cal H}=L^2(\hbox{\BB C})$ and let ${\cal H}_0$ be the eigenspace corresponding to the lowest eigenvalue (first Landau level) of the kinetic part of the Hamiltonian defined in (\ref{opb:fullham}) and let $P_0$ be the orthogonal projection onto ${\cal H}_0$. The Hamiltonian for our model is the operator on ${\cal H}_0$ given formally by \begin{equation} H\(\omega,\zeta\) = {\pi\over {2\kappa} } P_0 V\( \cdot , (\omega,\zeta)\) = {\pi\over {2\kappa} }P_0 V\( \cdot , (\omega,\zeta)\)P_0 \end{equation} where $(\omega,\zeta) \in \Omega$ and \begin{equation} V\(z, (\omega, \zeta)\) = \sum_{n\in\hbox{\BB Z} [i]} \omega_n \delta(z-\zeta_n). \end{equation} Note that $H$ coincides with $H_0$ in (\ref{opb:projham}) up to the term $BP_0$ and a multiplicative constant and that the lowest Landau energy is now shifted to zero. The projection $P_0$ is an integral operator with kernel \begin{equation}\label{opb:kernelproj} P_0 \(z,z'\) = {{2\kappa} \over { \pi}} \exp [- {\kappa} \vert z - z' \vert^2 - 2i \kappa z \wedge z'], \end{equation} where $\kappa=B/4$ and $ z \wedge z'={\cal R}z{\cal I}z'-{\cal I}z{\cal R}z'$, ${\cal R}z$ and ${\cal I}z$ being the real and imaginary parts of $z$ respectively. Note that if $\psi \in {\cal H}$ then $\psi \in {\cal H}_0$ if and only if $\psi(z) = f(z) e^{- \kappa \vert z \vert^2}$ where $f(z)$ is entire. Using (\ref{opb:kernelproj}) we can write the Hamiltonian in the form $$ H = \sum_{n\in\hbox{\BB Z} [i]} \omega_n f_{\zeta_n} \otimes {\overline{f_{\zeta_n}}}, $$ where for $\zeta \in \hbox{\BB C}$, \begin{equation} f_\zeta (z) = \sqrt{{\pi \over {2 \kappa}}}P_0 (z, \zeta) = \sqrt{{2\kappa} \over { \pi}} \exp [2\kappa {\bar \zeta}z- {\kappa}( \vert \zeta \vert^2 + \vert z \vert^2)]. \end{equation} Note that $\Vert f_\zeta \Vert =1$, $\langle f_\zeta\,,\ f_{\zeta'}\rangle =\sqrt{{\pi\over{2\kappa}}}f_{\zeta'}(\zeta)$ and that $H$ is an integral operator with kernel \begin{equation} H(z,z') = \sum_{n\in\hbox{\BB Z} [i]} \omega_n f_{\zeta_n} (z) {\overline{f_{\zeta_n} (z')}}. \end{equation} We first obtain a bound on $H(z,z^\prime )$ which implies that $H$ is bounded. We give the following simple estimate without proof. \noindent {\bf Lemma 2.1}:\ {\sl For $s,\ t > 0$ and $z,\ z'\in \hbox{\BB C}$ $$ \sum_{n \in \hbox{\BB Z}[i]} e^{-s \vert z -\zeta_n \vert^2} e^{-t \vert\zeta_n - z' \vert^2} \leq K(s+t) e^{- {{st} \over {s + t}} \vert z - z' \vert^2}. $$ where $$ K(s)= 9+8e^{-s} + 4\left( {\pi\over {s}}\right)^{1\over 2} +{4\over s} . $$ } The above Lemma implies that $\vert H(z,z')\vert$ is bounded above by \begin{equation} {{2R\kappa}\over \pi} K(2\kappa)e^{- {\kappa \over 2} \vert z - z' \vert^2}, \end{equation} where $R=\max(\vert a\vert, \vert b \vert)$. Therefore $H$ is bounded and \begin{equation} \Vert H \Vert \leq 4R K(2\kappa). \end{equation} Note that the heat kernel is $$ P_t(z,z')= {1\over{2\pi t}} e^{- {1 \over {2t}} \vert z - z' \vert^2} $$ and the corresponding operator has unit norm. From now on we take $\kappa$ sufficiently large so that $K(2\kappa )<10$ and we let ${\bar R}=40R$ so that $\Vert H\Vert\le {\bar R}$. \section{The Spectrum of $H$.} \paragraph Let $\{U_z:z\in\hbox{\BB C}\}$ be the family of unitary operators on ${\cal H}$ corresponding to the magnetic translations: $$ (U_zf)(z')= e^{2i\kappa z\wedge z'}f(z+z'). $$ Then for $m\in\hbox{\BB Z} [i]$ \begin{equation}\label{opb:magtrans} U_mH(\omega,\zeta)U_m^{-1}=H(\tau_m(\omega,\zeta)). \end{equation} Note that $[P_0,U_z]=0$ for all $z\in\hbox{\BB C}$ so that $U_z{\cal H}_0\subset {\cal H}_0$. Also $U_{z_1}U_{z_2} =e^{2i\kappa z_2\wedge z_1}U_{z_1+z_2}$. The ergodicity of $\{\tau_m:m\in\hbox{\BB Z} [i]\}$ and equation (\ref{opb:magtrans}) together imply that the spectrum of $H(\omega,\zeta)$ and its components are non random (see Ref. 14 Th V.2.4). {\bf Lemma 3.1}:\ {\sl With probability one} $$ [4a,4b]\subset\sigma \(H(\omega,\zeta)\). $$ {\bf Proof: } It is sufficient to prove that for each $E\in [4a,4b]$ and for all $\delta>0$, there exists $\Omega'\subset\Omega$ with $\hbox{\BB P} (\Omega')>0$ and $\psi\in {\cal H}_0$ with $||\psi||=1$ such that for all $(\omega,\zeta)\in\Omega'$, $||\(H(\omega,\ \zeta)-E\)\psi||<\delta.$ \vskip .2cm \noindent Let $B =\{0,1,i,1+i\}$. Choose $E\in [4a,4b]$ and $D$ such that $\sum_{n:\vert\zeta_n\vert\ge D}e^{- \kappa\vert\zeta_n -\zeta_0\vert^2}<\delta /4R$, where $R=\max (\vert a\vert ,\vert b\vert )$, and let $$ \Omega_2^\prime =\{\zeta\in\Omega_2 :\vert\zeta_n -\zeta_0\vert\le {\delta\over {4E\sqrt{2\kappa}}}\ \forall\, n\in B \} $$ then since for all $n\in B$, the impurities $\zeta_n$ and $\zeta_0$ can be arbitrarily close to one another, $\hbox{\BB P} (\Omega_2^\prime )>0$. Let $$ \Omega_1^\prime =\{\omega\in\Omega_1 :\vert\omega_n -{E\over 4}\vert <{\delta\over 16}\ \forall\, n\in B {\hbox {\rm { and }}} \max_{m\notin B\,:\vert\zeta_m\vert<D}\vert\omega_m \vert K(\kappa )<{\delta\over 4}\}. $$ Since $E/4$ and $0$ are in the support of $\mu$, $\hbox{\BB P} (\Omega_1^\prime )>0$. Let $\Omega^\prime =\Omega_1^\prime\times\Omega_2^\prime$, then $\hbox{\BB P} (\Omega^\prime )>0$. Now \begin{eqnarray*} (Hf_{\zeta_0}-Ef_{\zeta_0})(z) &=&\sum_{n\in B}(\omega_n -{E\over 4}) \langle f_{\zeta_n},f_{\zeta_0} \rangle f_{\zeta_n}(z)+E\left( {1\over 4}\sum_{n\in B} \langle f_{\zeta_n},f_{\zeta_0} \rangle f_{\zeta_n}(z)-f_{\zeta_0}(z)\right)\nonumber\\ &+&\sum_{n\notin B:\vert\zeta_n\vert <D}\omega_n \langle f_{\zeta_n}, f_{\zeta_0} \rangle f_{\zeta_n}(z)+\sum_{n:\vert\zeta_n\vert\ge D} \omega_n \langle f_{\zeta_n},f_{\zeta_0} \rangle f_{\zeta_n}(z). \end{eqnarray*} Hence \begin{eqnarray*} \Vert Hf_{\zeta_0}-Ef_{\zeta_0} \Vert &\le &{\delta\over 4}+{E\over 4} \Vert \sum_{n\in B} \langle f_{\zeta_n},f_{\zeta_0} \rangle f_{\zeta_n}-4f_{\zeta_0}\Vert\nonumber\\ &+&\sum_{n\notin B\,:\vert\zeta_n\vert <D}\vert\omega_n\vert\, \vert\langle f_{\zeta_n},f_{\zeta_0}\rangle \vert +R\sum_{n:\vert\zeta_n\vert\ge D} e^{-\kappa\vert\zeta_n -\zeta_0\vert^2}. \end{eqnarray*} It is easily seen that $\Vert \langle f_{\zeta_n},f_{\zeta_0}\rangle f_{\zeta_n}-f_{\zeta_0} \Vert^2 =1-\vert \langle f_{\zeta_n},f_{\zeta_0}\rangle \vert^2$ is bounded by $2\kappa\vert\zeta_n -\zeta_0\vert^2$, and therefore for all $(\omega ,\zeta )\in\Omega^\prime$, $$ \Vert Hf_{\zeta_0}-Ef_{\zeta_0} \Vert < \delta. $$ \par\noindent\rightline{$\square$} We now state the main theorem, which we will prove in the sequel. {\bf Theorem 3.2 } {\sl There exists $\kappa_0>0$ such that for $\kappa>\kappa_0$, with probability one, (a) $0$ is an eigenvalue of $H$ with infinite multiplicity, (b)\ $\sigma_{\hbox {\rm cont }}(H) =\emptyset$, (c) if $\lambda\in\sigma(H)\backslash\{ 0\}$, is an eigenvalue of $H$ with eigenfunction $\psi$, then $\psi$ decays exponentially with rate $\ge \kappa^{1/4}$. } We will start by showing part (a). The lemma is very close to results proven in Refs 8 and 13 so only a sketch of the proof will be given. {\bf Lemma 3.3}:\ {\sl There exists $\kappa_1 >0$ such that for $\kappa >\kappa_1$, with probability one, 0 is an eigenvalue of $H$ with infinite multiplicity.} \newline {\bf Proof: } Let $$ \psi_\zeta (z) = \prod_{n\in\hbox{\BB Z} [i]} (1 - {z \over \zeta_n}) e^{{z\over \zeta_n} +{{z^2}\over {2\zeta_n^2}}}. $$ If we can show that the sums $\sum_n 1/\vert\zeta_n\vert^3$ and $\sum_n 1/\zeta_n^2$ converge independently of $\zeta$ then it follows from the theory of entire functions (see Ref. 15) that there exists $A>0$ and $R>0$, both independent of $\zeta$ such that for $\vert z\vert>R$, $\vert \psi_\zeta (z)\vert \leq e^{A\vert z \vert^2}$. The first sum is easily bounded, the second can be bounded by utilising the four-fold rotational symmetry of the $\hbox{\BB Z} [i]$ to cancel any large contributions. Let $\phi_k (z) = z^k \psi_\zeta (z) e^{-\kappa \vert z \vert^2}$ for $k \geq 1$, then if $\kappa >A$, the $\phi_k$'s are in ${\cal H}_0$ and $\phi_k (\zeta_n)=0$ for all $n\in\hbox{\BB Z} [i]$. Therefore $H\phi_k =0$ for all $k\ge 1$. Moreover if $\sum^N_{j=1} a_{k_j} \phi_{k_j} = 0$ then $\sum^N_{j=1} a_{k_j} z^{k_j} = 0$ for $ z \notin \{ \zeta_n \}$. Therefore $\sum^N_{j=1} a_{k_j} z^{k_j} \equiv 0$ and thus the $ a_{k_j}$'s are zero implying that the $\phi_k$'s are independent.\par\noindent\rightline{$\square$} For Theorem 3.2 parts (b), (c) we can simplify the problem by studying the behaviour of the generalized eigenfunctions at the impurity sites only. We have $$ (H\psi )(z)={\pi\over {2\kappa}}\sum_{n\in\hbox{\BB Z} [i]}\omega_n P_0 (z,\zeta_n )\psi (\zeta_n ) $$ and thus if $(H\psi )(z)=\lambda\psi (z)$, \begin{equation}\label{opb:sitetise} {\pi\over {2\kappa}}\sum_{n\in\hbox{\BB Z} [i]}\omega_n P_0 (z,\zeta_n )\psi (\zeta_n )=\lambda\psi (z) \end{equation} which evaluated at $\zeta_m$ gives $$ {\pi\over {2\kappa}}\sum_{n\in\hbox{\BB Z} [i]}\omega_n P_0 (\zeta_m ,\zeta_n ) \psi (\zeta_n )=\lambda \psi (\zeta_m ) $$ Let $\omega_n \psi (\zeta_n )=\xi_n$. Then \begin{equation} {\pi\over {2\kappa}}\sum_{n\in\hbox{\BB Z} [i]} P_0 (\zeta_m ,\zeta_n ) \xi_n ={\lambda\over {\omega_m}}\xi_m \end{equation} We can thus reduce the problem to the study of a random matrix which has $\omega$-dependent elements on the diagonal and $\zeta$-dependent rapidly decaying off-diagonal elements. We write this matrix as a sum of a diagonal matrix and an off-diagonal matrix as defined below. Let ${\cal M}=l^2 (\hbox{\BB Z} [i]) $. Define the operators $M_0$, and $V_{\omega}^{\lambda}$ on ${\cal M}$ as follows. \begin{eqnarray}\label{opb:defnM} & &\langle m\vert M_0\vert n\rangle ={\pi\over 2\kappa}P_0(\zeta_m ,\zeta_n)(1-\delta_{mn})\nonumber\\ & &\langle m\vert V_{\omega}^{\lambda}\vert n\rangle =\left( 1-{\lambda\over \omega_n}\right) \delta_{mn}. \end{eqnarray} For a proof of Theorem 3.2 part (c) we note that the eigenvectors $\xi$ of $M^{\lambda}=M_0+V_{\omega}^{\lambda}$, are related by an explicit formula to the generalized eigenfunctions $\psi$ of $H$ in such a way that exponential decay of the former implies exponential decay of the latter: >From (\ref{opb:sitetise}), if $\lambda\ne 0$ $$ \psi (z)={\pi\over {2\kappa\lambda}}\sum_{n\in\hbox{\BB Z} [i]} P_0(z,\zeta_n )\xi_n $$ If $\xi_n $ decays exponentially, $\vert\xi_n \vert\le Ce^{-m\vert\zeta_n \vert }$ we have \begin{eqnarray}\label{opb:expimpliesexp} \vert\psi (z)\vert &\le &{C\over\lambda}\sum_{n\in\hbox{\BB Z} [i]} e^{-\kappa\vert z-\zeta_n\vert^2}e^{-m\vert\zeta_n \vert} \le {C\over\lambda }e^{-m\vert z\vert}\sum_{n\in\hbox{\BB Z} [i]} e^{-\kappa\vert z-\zeta_n\vert^2}e^{m\vert z-\zeta_n \vert}\nonumber\\ &\le & {C\over\lambda}e^{-m\vert z\vert} e^{{m^2}\over {2\kappa }} \sum_{n\in\hbox{\BB Z} [i]} e^{-{\kappa\over 2}\vert z-\zeta_n\vert^2} \le {C\over\lambda}e^{{m^2}\over {2\kappa }}K({\kappa\over 2}) e^{-m\vert z\vert} \end{eqnarray} and $\psi (z)$ decays exponentially, where we have bounded the sum by taking $s=\kappa /2$, $t=0$ in Lemma 2.1. Thus we want to show that the eigenvectors for the eigenvalue equation $M^{\lambda}\xi =0$ decay exponentially for $\lambda\ne 0$. We will do this by the same method as in Ref. 8. First we will need to make a few definitions. For regions $\Lambda \subset \hbox{\BB Z}[i]$ we define $M^\lambda_\Lambda$ to be the restriction of $M^\lambda $ to $l^2(\Lambda)$. If $E \notin \sigma (M^\lambda_\Lambda)$ then the Green function \begin{equation} \Gamma^\lambda_\Lambda(E) = (M^\lambda_\Lambda - E)^{-1} \end{equation} is well-defined. In particular, we shall consider the regions \begin{equation} \Lambda_L (n) = \{ n' \in \hbox{\BB Z}[i] \vert\ :\ \vert n'-n \vert_\infty <L/2 \} \end{equation} for $n \in \hbox{\BB Z}[i]$ and $L > 0$. {\bf Definition. }{\it Fix constants $\beta \in (0,1)$ and $s \in ({1\over 2},1)$. Given a configuration $(\omega ,\zeta )$, a square $\Lambda_L (n)$ is called $(m,E)$-regular for some $m>0$ and $E \in \hbox{\BB R}$ if the following two conditions are satisfied: $$ d(E,\sigma(M^\lambda_{\Lambda_L(n)}(\omega ,\zeta ))) >{1\over 2} e^{-L^\beta}, \leqno (RA) $$ $$ \vert\langle n \vert \Gamma^\lambda_{\Lambda_L(n)}(E)\vert n'\rangle\vert\leq e^{-mL} \leqno (RB) $$ for all $n'\in \tilde \Lambda_L (n) $ where $\tilde \Lambda_L (n) = \Lambda_L(n) \setminus \Lambda_{\tilde L} (n) $ with $\tilde L = L - L^s$. $\Lambda_L (n)$ is called singular if it is not regular.} We now state a theorem which is an variant of the main theorem in Ref. 2 where the von Dreifus and Klein Theorem is adapted from the case of a tight-binding (finite range) Hamiltonian to the case where the Hamiltonian has a long range hopping term with Gaussian decay. It states conditions under which the eigenvectors of the random matrix $M^\lambda$ with eigenvalue $0$ decay exponentially. {\bf Theorem 3.4} {\it Fix constants $\beta \in (0,1)$, $s \in ({1\over 2},1)$, $\gamma\in (0,1)$, $p > 2$, $q > 4p+12$. There exists $Q_0 > 0$ depending on all these constants but {\bf independent of $\lambda$ and $\kappa >1$} such that the following holds: If for $\lambda,\,\kappa$ the conditions {\rm (P1)} and {\rm (P2)} are satisfied, where \begin{description} \item[\hspace{0.5cm}] {\rm (P1)} There exists an $L_0 > Q_0$ and $m_0 $ such that \begin{equation} \hbox{\BB P} \left\{ \Lambda_{L_0} (0) \hbox{ is } (m_0,0)\hbox{-regular} \right\} \geq 1-L_0^{-p} \end{equation} \item[\hspace{0.5cm}] {\rm (P2)} There exists $\eta >0$ such that, for all $E \in (-\eta,\eta)$ and for all $L > L_0$, \begin{equation} \hbox{\BB P} \left\{ d\left(E , \sigma \left( M^\lambda_{\Lambda_L (0)} \right) \right) < e^{-L^\beta} \right\} < L^{-q}, \end{equation} \end{description} \noindent then, for all $m \in (0,m_0)$, there exists $\delta>0$ depending on $m,m_0,L_0,\beta$ and $\eta$ such that for all $E\in (-\delta,\delta)$ the eigenvectors of $M^\lambda$ with eigenvalue $E$ decay exponentially with rate $\geq m$.} The main work of this paper consists in proving that the conditions (P1) and (P2) are satisfied. The conditions can be seen to consist of two types of estimate. $(RB)$ of (P1) is an estimate of the decay of the Greens function $\Gamma^\lambda_{\Lambda_L}(0)$, while $(RA)$ of (P1) and (P2) are Wegner type estimates that require small gaps in the $\Lambda_L$ dependent spectrum. It is unusual that it is the latter estimates that will require the finer analysis; previous works have found the decay of the Green's function to require the more delicate study. This is because we want to show that the eigenfunctions are exponentially decaying for arbitrarily small $\lambda$. Inspecting (\ref{opb:defnM}) we see that for $\lambda$ small the $\omega$-dependence becomes less significant and does not give sufficient randomness for Wegner type estimates. Therefore we have to utilise randomness provided by the positions of the $\zeta_j$'s. \section{Proof of the Conditions (P1) and (P2)} \paragraph We will begin by showing $(RB)$ of (P1). We will need the following two probabilistic lemmas in which we fix $u>3$. {\bf Lemma 4.1}: {\sl There exists $Q_1$ such that \begin{equation} \hbox{\BB P}\left(\vert\zeta_n -\zeta_{n'}\vert>{2\over L^u} {\hbox { for all }} n,n'\in\Lambda_L,\ n\ne n' \right) \ge 1-{1\over L^{u-3}} \end{equation} for all $L>Q_1$. } {\bf Proof}: The $\zeta_n$'s have a distribution with a density bounded by $r_b$ for each $B_n$ and thus, $$ \hbox{\BB P}\left(\vert\zeta_n -\zeta_{n'}\vert>{2\over L^u} {\hbox { for all }} n,n'\in\Lambda_L,\ n\ne n' \right) \ge\left(1-{{4r_b}\over L^u}\right)^{L^2}. $$ By taking $L$ sufficiently large we get the result.\par\noindent\rightline{$\square$} {\bf Lemma 4.2}: {\sl There exists $Q_2$ such that \begin{equation} \hbox{\BB P}\left(\left\vert 1-{\lambda\over\omega_n}\right\vert >{1\over L^u} \ \ \forall\ n\in\Lambda_L\ \right)> 1-{1\over L^{u-3}} \end{equation} for all $L>Q_2$. } {\bf Proof}: Now $\displaystyle {\left\vert 1-{\lambda\over\omega_n}\right\vert\le {1\over L^u}}$ gives us that $\displaystyle {-{1\over L^u}\le 1-{\lambda\over\omega_n} \le {1\over L^u}}$ which implies \newline that $\displaystyle {1+{1\over L^u}\ge {\lambda\over\omega_n} \ge 1-{1\over L^u}}$. Thus $\omega_n$ must fall between the bounds, \begin{equation} {{\vert\lambda\vert}\over {1+{1\over L^u}}} \le\ \vert\omega_n\vert\ \le{{\vert\lambda\vert}\over {1-{1\over L^u}}} . \end{equation} Hence \begin{eqnarray*} \hbox{\BB P}\left(\left\vert 1-{\lambda\over\omega_n}\right\vert\le {1\over L^u} \right)&\le &2\rho_b \vert\lambda\vert\left({1\over {1- {1\over L^u}}}- {1\over {1+ {1\over L^u}}}\right) \nonumber\\ &=&{{4\rho_b\vert\lambda\vert}\over L^u}(1-{1\over L^{2u}})^{-1} \le {{2^{u+2}\rho_b {\bar R}}\over L^u}. \end{eqnarray*} if $L>2$, where we have used $1-1/L^{2u}>1/2^u$. Therefore $$ \hbox{\BB P}\left(\left\vert 1-{\lambda\over\omega_n}\right\vert >{1\over L^u} \ \ \forall\ n\in\Lambda_L\ \right)>\left( 1- {{2^{u+2} \rho_b {\bar R}}\over L^u} \right)^{(L+1)^2}. $$ By taking $L$ sufficiently large we get the result.\par\noindent\rightline{$\square$} The following Lemma is proved in Ref. 8. {\bf Lemma 4.3}:\ {\sl For all $\gamma\in (0,1)$, there exists $C_0(\gamma)> 0$ such that for $\alpha>1$} \begin{equation} \sum_{m \in \hbox{\BB Z}[i]} e^{- \alpha \{ \vert z - m \vert^\gamma + \vert z' - m \vert^\gamma\}} \leq C_0(\gamma) e^{- \alpha \vert z - z' \vert^\gamma}. \end{equation} The following Lemma will be used to show $(RB)$ of (P1). {\bf Lemma 4.4}: {\sl For all $\gamma\in (0,1)$ and $u>3$, there exists $Q_3$ such that for all $L> Q_3$ and all $\kappa >L^{4u}/4$ we have for any $n,n'\in\Lambda_L$, \begin{equation} \hbox{\BB P}\(\vert\langle n\vert\Gamma_{\Lambda_L}^{\lambda}(0)\vert n'\rangle\vert\le 2L^u e^{-{\kappa^{1/2}\over 8}\vert n-n'\vert^{\gamma}}\)>1-{2\over L^{u-3}} . \end{equation} } {\bf Proof}: In the following we let $\gamma\in (0,1)$. Using $\vert a+b\vert^2\le 2(\vert a\vert^2+\vert b\vert^2 )$ we have $\kappa\vert\zeta_n -\zeta_{n'}\vert^2\ge\kappa({1\over 2}\vert n-n'\vert^2 -2) >{\kappa\over 4}\vert n-n'\vert^2>{{\kappa^{1/2}}\over 4} \vert n-n'\vert^{\gamma}$ for $\vert n-n'\vert\ge 3$. Suppose that $\vert\zeta_n-\zeta_{n'}\vert>2/L^u$ for all $n,n'\in\Lambda_L$, $n\ne n'$. Then for $\vert n-n'\vert <3$ we have that $\kappa\vert\zeta_n -\zeta_{n'}\vert^2> {{\kappa^{1/2}}\over 4}\vert n-n'\vert^{\gamma}$ if ${{2\kappa^{1/2}}\over L^{2u}}> 1$. Thus we can write $$ e^{-\kappa\vert\zeta_n -\zeta_{n'}\vert^2} \le e^{-{\kappa^{1/2}\over 4}\vert n -n'\vert^{\gamma}} $$ and consequently $$ \vert\langle n\vert M_{\Lambda_L}^0\vert n'\rangle\vert \le e^{-{\kappa^{1/2}\over 4}\vert n -n'\vert^{\gamma}}(1-\delta_{n n'}). $$ If $\vert 1-\lambda/\omega_n\vert>1/L^u$ for all $n\in\Lambda_L$ then for all $n,n'\in\Lambda_L$ we also have, $$ \vert\langle n\vert (V_{\Lambda_L}^{\lambda} )^{-1}\vert n'\rangle\vert\le L^{u} \delta_{n n'}. $$ Therefore we can write \begin{eqnarray*} \vert\langle n\vert (V_{\Lambda}^{\lambda})^{-1} M_{\Lambda_L}^0\vert n'\rangle\vert\, && \le\,\sum_{p\in\Lambda}\vert\langle n\vert (V_{\Lambda}^{\lambda})^{-1}\vert p\rangle\vert\vert\langle p\vert M_{\Lambda_L}^0\vert n'\rangle\vert\\ &&=\,\vert\langle n\vert (V_{\Lambda}^{\lambda})^{-1}\vert n\rangle\vert\vert\langle n\vert M_{\Lambda_L}^0\vert n'\rangle\vert\\ &&\le\, L^u e^{-{\kappa^{1/2}\over 8}} e^{-{\kappa^{1/2}\over 8}\vert n -n'\vert^{\gamma}} (1-\delta_{nn'}), \end{eqnarray*} and \begin{eqnarray*} \vert\langle n\vert\left( (V_{\Lambda_L}^{\lambda})^{-1} M_{\Lambda_L}^0 \right)^2\vert n'\rangle\vert &&\le\ \sum_{r\in\Lambda}\vert\langle n\vert (V_{\Lambda_L}^{\lambda})^{-1}\vert n\rangle\vert\vert\langle n\vert M_{\Lambda_L}^0\vert r\rangle\vert \vert\langle r\vert (V_{\Lambda_L}^{\lambda})^{-1}\vert r\rangle\vert\vert\langle r\vert M_{\Lambda_L}^0\vert n'\rangle\vert\\ &&\le\ L^{2u}e^{-{\kappa^{1/2}\over 4}} \sum_{r\in\hbox{\BB Z} [i]}e^{-{\kappa^{1/2}\over 8}\vert n-r\vert^{\gamma}} e^{-{\kappa^{1/2}\over 8}\vert r-n'\vert^{\gamma}} (1-\delta_{nr})(1-\delta_{rn'})\\ &&\le\ C_0(\gamma)L^{2u}e^{-{\kappa^{1/2}\over 4}} e^{-{\kappa^{1/2}\over 8}\vert n-n'\vert^{\gamma}}. \end{eqnarray*} Similarly, $$ \vert\langle n\vert\left( (V_{\Lambda_L}^{\lambda})^{-1} M_{\Lambda_L}^0\right)^k \vert n'\rangle\vert \le C_0(\gamma)^{k-1}L^{ku}e^{-{{k\kappa}\over 8}^{1/2}} e^{-{\kappa^{1/2}\over 8}\vert n-n'\vert^{\gamma}}. $$ Let $T$ be the operator with $\langle n\vert T\vert m \rangle =e^{-{\kappa^{1/2}\over 8}\vert n-m\vert^{\gamma}}$. Then we can make $\Vert\left( (V_{\Lambda_L}^{\lambda})^{-1} M_{\Lambda_L}^0\right)^k\Vert\le {1\over 2^k}\Vert T\Vert$ by making $C_0(\gamma ) L^u e^{-{\kappa^{1/2}\over 8}}<{1\over 2}$. We can therefore iterate the resolvent identity to get, \begin{eqnarray*} \Gamma^{\lambda}(0)&=&(V^{\lambda})^{-1}-(V^{\lambda})^{-1}M^0\Gamma^{\lambda}(0)\nonumber\\ &=&\sum_{k=0}^{\infty}(-1)^k \left((V^{\lambda})^{-1}M^0\right)^k (V^{\lambda})^{-1} \end{eqnarray*} Hence, if we take $L>Q_3$ with $Q_3$ larger than $Q_1$ and $Q_2$ and sufficiently large that ${1\over 2} L^{2u}>8\ln (2C_0(\gamma )L^u)$, the result follows from Lemmas 4.1 and 4.2.\par\noindent\rightline{$\square$} \paragraph Let $\beta$ be fixed as in Theorem 3.4 and $\kappa>\pi /2$. To prove $(RA)$ of (P1) and condition (P2) we need to look at two regimes, $\vert\lambda\vert\ge e^{-L^\beta/2}$ and $\vert\lambda\vert <e^{-L^\beta/2}$. The next lemma deals with the first regime, and the Lemmas 4.6 - 4.9 with the second. {\bf Lemma 4.5}: {\sl If $\vert\lambda\vert\ge e^{-L^\beta/2}$, $L>1$, $E\in\hbox{\BB R}$ and $\epsilon >0$, then \begin{equation} \hbox{\BB P} ( d(E, \sigma (M^{\lambda}_{\Lambda_L})) <\epsilon ) <8\rho_b R^2L^2 e^{L^\beta/2} \epsilon . \end{equation} } {\bf Proof}: First we need to find a bound on the density of the diagonal terms of $M^{\lambda}$. \begin{eqnarray}\label{opb:densitybound} \sup_x \lim_{\epsilon\rightarrow 0}{1\over {2\epsilon}} \int_{x-\epsilon <1-{\lambda\over \omega}<x+\epsilon}\rho (\omega )d\omega &=&\sup_x \lim_{\epsilon\rightarrow 0}{1\over {2\epsilon\lambda}} \int_{x-\epsilon}^{x+\epsilon} \rho\left( {\lambda\over {1-u}}\right)\left( {\lambda\over {1-u}}\right)^2 du\nonumber\\ &<&\sup_x \lim_{\epsilon\rightarrow 0} {{\rho_b R^2}\over {2\epsilon\vert\lambda\vert}} \int_{x-\epsilon}^{x+\epsilon}du ={{\rho_b R^2}\over \vert\lambda\vert}. \end{eqnarray} It follows that the density of $x_{nn}=\langle n\vert M_{\Lambda}^{\lambda}\vert n\rangle$ is bounded by $\rho_b R^2 e^{L^\beta/2}$. For Borel subsets $B$ of $\hbox{\BB R}$ let $\sigma^\Lambda_n(B) = \langle n \vert E_\Lambda (B) \vert n \rangle$, where $ E_\Lambda (B) $ are the spectral projections of $ M^{\lambda}_\Lambda $. Then by Lemma VIII.1.8 in Ref. 14, and by (\ref{opb:densitybound}) $$ \hbox{\BB E}_{x_{nn}} \sigma^\Lambda_n (B) < \rho_b R^2 e^{L^\beta/2} \int_B dx $$ and therefore $$ \hbox{\BB E} \sigma^\Lambda_n (B) < \rho_b R^2 e^{L^\beta/2}\int_B dx. $$ As in Proposition VIII.4.11 of Ref. 14, it then follows that, using (\ref{opb:densitybound}), for all $E\in\hbox{\BB R}$ and $\epsilon >0$, $$ \hbox{\BB P} ( d(E, \sigma (M^{\lambda}_{\Lambda_L})) <\epsilon ) <2\rho_b R^2 e^{L^\beta/2} \epsilon \vert \Lambda_L \vert <2\rho_b R^2 e^{L^\beta/2} \epsilon (L+1)^2. $$ If $L\ge 1$ the result follows.\par\noindent\rightline{$\square$} For the next part it is necessary to make some definitions. {\bf Definitions:} We define $V=\{\zeta_n\, , n\in\hbox{\BB Z} [i]\}$ to be the vertices of a graph ${\cal G}(V,E)$ with edges $E$ defined as $E=\{ (\zeta_m,\zeta_n)\ :\ \vert\zeta_m-\zeta_n\vert <1/8,\ \forall\ n,m \in \hbox{\BB Z} [i], n\ne m \}$. The {\sl degree} of a vertex, ${\hbox {\rm deg}}(\zeta_m)=\#\{n\in\hbox{\BB Z} [i]:\ (\zeta_m,\zeta_n)\in E\}$. Two vertices are said to be {\sl connected} if there exists a path between them along a series of edges. A {\sl component} is defined to be a maximally connected subgraph. We will define a {\sl cluster} to be a component of the graph ${\cal G}(V,E)$. \par {\bf Lemma 4.6}: {\sl For each configuration $\{\zeta_n\}$ there exist clusters containing at most 4 vertices such that the distance between every pair of clusters is at least $1/8$.} {\bf Proof}: The distance between two clusters ${\cal C}_i$, ${\cal C}_j$ is given by, $$ d({\cal C}_i,{\cal C}_j)=\inf\{\, d(\zeta_n^i,\zeta_m^j)\,\vert\,\zeta_n^i\in {\cal C}_i,\,\zeta_m^j\in {\cal C}_j\}. $$ It is easily seen that if the distance between two clusters is less than $1/8$, then the distance between one of the vertices in one cluster, must be closer than $1/8$ to a vertex in the other. Thus an edge will exist that connects the two clusters, leading to a contradiction in their definition as separate clusters. It suffices to show that we cannot have a cluster with more than four vertices. The diameter of a cluster is given by, $$ {\hbox {\rm diam}}({\cal C}_i)=\sup\{\, d(\zeta_n,\zeta_m)\ \vert\ \zeta_n,\zeta_m\in {\cal C}_i\}. $$ We know that the unit squares centred on the Gaussian integers, $\{B_n, n\in\hbox{\BB Z} [i]\}$, contain exactly one vertex each. The maximum diameter for a cluster of five vertices will be less than 1/2 by virtue of the definition of a cluster. However a circle of diameter 1/2 cannot intersect more than four of the $B_n$, so we cannot have a cluster of five. If we had a cluster of more than five vertices, we could also have a cluster of five as can be seen if we perform a one by one deletion of lowest degree vertices until only five remain. Thus we cannot have a cluster with more than five vertices.\par\noindent\rightline{$\square$} For a configuration $\{\zeta_n\}$, let $$ \langle n\vert {\tilde M}^{\lambda}_{\Lambda}\vert n'\rangle = \cases{\langle n\vert M^{\lambda}_{\Lambda}\vert n'\rangle & if $\zeta_n ,\zeta_{n'}$ are in the same cluster, \cr {0} & otherwise. \cr} $$ Let ${\cal C}_1, {\cal C}_2,\ldots ,{\cal C}_N$ be the clusters in $\Lambda$ and let ${\cal P}_1, {\cal P}_2, \ldots ,{\cal P}_N$ be the projections onto ${\cal H}_i$ the space spanned by $\{\vert n\rangle\, :\,\zeta_n\in {\cal C}_i\}$. Let \begin{equation}\label{opb:M_i} M_i={\cal P}_i M^{\lambda}_{\Lambda} {\cal P}_i ={\cal P}_i {\tilde M}^{\lambda}_{\Lambda} {\cal P}_i. \end{equation} {\bf Lemma 4.7}: {\sl For $\lambda =0$, \begin{equation} \vert\det M_i\vert\ge\quad C\prod_{m<\, n :\, \zeta_m ,\,\zeta_n\in\, {\cal C}_i} \(1-e^{-\kappa\vert\zeta_m -\zeta_n\vert^2}\), \end{equation} where $C>0$ is a constant independent of $\zeta$. } Note that numerical calculation shows that the inequality is satisfied with $C=1$. If the $\zeta_i$'s are distinct then for $\lambda =0$ we can write \begin{eqnarray}\label{lemm_ineq} \langle \xi, M_i\xi\rangle &=&{\pi\over 2\kappa}\sum_{m,n:\zeta_m,\zeta_n\in {\cal C}_i}{\bar \xi}_m P_0(\zeta_m,\zeta_n)\xi_n={\pi\over 2\kappa}\int_{\hbox{\BB C}}\sum_{m,n:\zeta_m,\zeta_n\in {\cal C}_i}{\bar \xi}_m P_0(\zeta_m,z)P_0(z,\zeta_n)\xi_n\ dz\nonumber\\ &=& \int_{\hbox{\BB C}}\vert\sum_{m:\zeta_m\in {\cal C}_i}\xi_m f_{\zeta_m}(z)\vert^2dz\ >\ 0 \end{eqnarray} since the $f_{\zeta_i}$'s will be linearly independent. Thus $\det M_i>0$. {\bf Proof:} If $\lambda =0$, recall that from the definition of $M^{\lambda}$, for $\zeta_m$, $\zeta_n$ in a cluster ${\cal C}_i$, $$ \langle m\vert M_i\vert n\rangle =e^{-\kappa\vert\zeta_m-\zeta_n\vert^2-2i\kappa\zeta_m \wedge\zeta_n}. $$ We only have to prove the result up to a cluster of four. For a cluster of one, the result is trivial. For a cluster of two we get, $$ \vert\det M_i\vert =1-e^{-2\kappa\vert\zeta_1-\zeta_2\vert^2}. $$ We now give the proof for a cluster of three. A direct proof with $C=1$ can be given (see Ref 16) but it is difficult to extend this to the case of a cluster of four impurities. For this reason we shall give an indirect proof which can be extended to the latter case. Let $\kappa^{{1\over 2}}(\zeta_2-\zeta_1)=a e^{i\alpha}$ and $\kappa^{{1\over 2}}(\zeta_3-\zeta_1) =b e^{i\beta}$. Then $\det M_i = G_3(a,b,\phi)$ where $\phi =\alpha -\beta$ and \begin{equation}\label{main1} G_3(a,b,\phi)=1 - e^{-2a^2}- e^{-2b^2}-e^{-2c^2}+2e^{-(a^2+b^2+c^2)}\cos(2ab\sin (\phi)), \end{equation} with $$ c^2=a^2+b^2-2ab\cos\phi. $$ Note that without loss of generality we can take $\phi\in [0,\pi ]$. $G_3$ is an analytic function of $a$, $b$ and $\phi$. It is easy to check that, $$ G_3(0,b,\phi)=G_3(a,0,\phi)=G_3(a,ae^{\pm i\phi},\phi)=0, $$ and $$ \frac{\partial G_3} {\partial a }(0,b,\phi) =\frac{\partial G_3} {\partial b }(a,0,\phi)=0 $$ so that we can write \begin{equation}\label{factor1} G_3(a,b,\phi)=a^2b^2(b-ae^{-i\phi})(b-ae^{i\phi})g_3(a,b,\phi)=a^2b^2c^2g_3(a,b,\phi) \end{equation} where $g_3(a,b,\phi)$ is an analytic function of $a$, $b$ and $\phi$. Let $A=\dot{\hbox{\BB R}}_+^2\times [0,\pi ]$, where $\dot{\hbox{\BB R}}_+$ denotes the one-point compactification of $\hbox{\BB R}_+$, and let $A_o$ be the interior of $A$. Define $f_3:A_o\rightarrow\hbox{\BB R}$ by \begin{equation}\label{total1} f_3(a,b,\phi)={G_3(a,b,\phi)\over (1-e^{-a^2})(1-e^{-b^2})(1-e^{-c^2})}. \end{equation} $f_3(a,b,\phi)>0$ for all $(a,b,\phi )\in A_o$ by the inequality (\ref{lemm_ineq}). Note that $c=0$ only if $a=b=0$ or $a=b$ and $\phi =0$. We shall prove that for each point $(a_0,b_0,\phi_0)$ on the boundary of $A$, we have $\displaystyle{\liminf_{(a,b,\phi)\rightarrow (a_0,b_0,\phi_0)}}$ $f_3(a,b,\phi)>0$. Then since $A$ is compact there exists $C>0$ such that $f_3(a,b,\phi)>C$ for all $(a,b,\phi)\in A$. For points on the boundary of $A$ for which $a$, $b$ and $c$ are all finite and non-zero $f_3(a,b,\phi )$ is defined by (\ref{total1}) and is strictly positive. Now $\lim_{a\rightarrow\infty}f_3(a,b,\phi)=1+e^{-b^2}>1$ for all $(b,\phi )\in\hbox{\BB R}_+\times [0,\pi ]$ and similarly for $\lim_{b\rightarrow\infty}f_3(a,b,\phi)$. Also $\displaystyle{\liminf_{(a,b)\rightarrow(\infty,\infty )}}$ $f_3(a,b,\phi)=1$. Next we have that $\lim_{(a,b)\rightarrow (0,0)}$ $f_3(a,b,\phi)=g_3(0,0,\phi)$ and we can check that $g_3(0,0,\phi )=4$. For $b>0$, $\lim_{a\rightarrow 0} f_3(a,b,\phi)=b^4g_3(0,b,\phi)/(1-e^{-b^2})^2$. We can calculate $g_3(0,b,\phi )$ explicitly to get $b^4g_3(0,b,\phi)=2e^{-2b^2}(e^{2b^2}-1-2b^2)>0$. Similarly we can show that $g_3(a,0,\phi )>0$. Finally, by symmetry it follows that $\lim_{a\rightarrow b,\ \phi\rightarrow 0}f_3(a,b,\phi)=\lim_{a\rightarrow 0}f_3(a,b,\psi)$ where $\psi$ is the angle between the edges of lengths $b$ and $c$, which has already been shown to be strictly positive. Note that in fact this limit is independent of $\psi$. Now we come to the proof of the Lemma for a cluster of four. The idea of the proof is the same as for a cluster of three but the details are more complicated. Let $\kappa^{{1\over 2}}(\zeta_2-\zeta_1)=a e^{i\alpha}$, $\kappa^{{1\over 2}}(\zeta_3-\zeta_1)=b e^{i\beta}$ and $\kappa^{{1\over 2}}(\zeta_4-\zeta_1)=c e^{i\gamma}$. Then $\det M_i = G_4(a,b,c,\phi ,\psi)$ where $\phi =\beta -\alpha $, $\psi =\alpha -\gamma$ and $$ G_4(a,b,c,\phi,\psi)=1 - e^{-2a^2}- e^{-2b^2}-e^{-2c^2} -e^{-2u^2} -e^{-2v^2}-e^{-2w^2} $$ $$ + e^{-2(b^2+v^2)}+e^{-2(a^2+w^2)}+e^{-2(c^2+u^2)} $$ $$ +2e^{-(a^2+b^2+u^2)}\cos(4\Delta_{abu}) +2e^{-(b^2+c^2+w^2)}\cos(4\Delta_{bcw}) $$ $$ +2e^{-(a^2+c^2+v^2)}\cos(4\Delta_{acv}) +2e^{-(u^2+v^2+w^2)}\cos(4\Delta_{uvw}) $$ $$ -2e^{-(b^2+c^2+u^2+v^2)}\cos(4(\Delta_{acv}+\Delta_{abu})) $$ $$ -2e^{-(a^2+ c^2+u^2+w^2)}\cos(4(\Delta_{abu}-\Delta_{bcw})) $$ \begin{equation}\label{main2} -2e^{-(a^2+b^2+v^2+w^2)}\cos(4(\Delta_{acv}-\Delta_{bcw})) \end{equation} with $$ u^2=b^2+a^2-2ba\cos\phi\ , $$ $$ v^2=c^2+a^2-2ac\cos\psi\ , $$ $$ w^2=b^2+c^2-2bc\cos(\phi+\psi)\ , $$ $$\Delta_{abu}={1\over 2} ba\sin\phi\ , $$ $$ \Delta_{acv}={1\over 2} ac\sin\psi\ , $$ $$ \Delta_{bcw}={1\over 2} bc\sin(\phi+\psi)\ , $$ and $$ \Delta_{uvw}={1\over 2}\left( ba\sin\phi+ac\sin\psi-bc\sin(\phi+\psi)\right)\ . $$ $G_4(a,b,c,\phi,\psi)$ is an analytic function of $ a,b,c $, $\phi$ and $\psi$. In this case also we can check that, $$ G_4(0,b,c,\phi,\psi)=G_4(a,0,c,\phi,\psi)=G_4(a,b,0,\phi,\psi)=0, $$ $$ \frac{\partial G_4} {\partial a }(0,b,c,\phi,\psi) =\frac{\partial G_4} {\partial b }(a,0,c,\phi,\psi) =\frac{\partial G_4} {\partial c }(a,b,0,\phi,\psi)=0, $$ and $$ G_4(be^{\pm i\phi},b,c,\phi,\psi)=G_4(ce^{\pm i \psi},b,c,\phi,\psi) =G_4(a,ce^{\pm i(\phi+\psi)}\,c,\phi,\psi)=0. $$ These identities imply that \begin{equation}\label{factor2} G_4(a,b,c,\phi,\psi)=a^2b^2c^2u^2v^2w^2 g(a,b,c,\phi,\psi) \end{equation} where $ g_4(a,b,c,\phi,\psi)$ is an analytic function of $ a,b,c $, $\phi$ and $\psi$. In this case we let $A=\dot{\hbox{\BB R}}_+^3\times [0,\pi ]^2$ and define $f_4:A_o\rightarrow \hbox{\BB R}$ by \begin{equation}\label{total2} f_4(a,b,c,\phi,\psi)={G_4(a,b,c,\phi,\psi)\over (1-e^{-a^2})(1-e^{-b^2})(1-e^{-c^2})(1-e^{-u^2})(1-e^{-v^2})(1-e^{-w^2})}. \end{equation} Using the same arguments as before it is sufficient to check that for each point $z=(a_0,b_0,c_0,\phi_0,\psi_0)$ on the boundary of $A$, we have $\displaystyle{\liminf_{\smash{(a,b,c,\phi,\psi)\rightarrow z}}}\, f_4(a,b,c,\phi,\psi)>0$. When one of $a$, $b$ and $c$ tend to $\infty$, the problem simplifies to the three impurity case and taking the lower limit when two of them tend to $\infty$, reduces the problem to the two impurity case. When all of $a$, $b$ and $c$ tend to $\infty$ the lower limit is equal to 1. It remains to show that $f_4(a,b,c,\phi ,\psi )$ is strictly positive in the limit of any subset of $\{a,b,c\}$ going to zero. By symmetry we need only check the cases $a,b,c\rightarrow 0$, $a,b\rightarrow 0$ and $a\rightarrow 0$. Now $\lim_{(a,b,c)\rightarrow (0,0,0)}$ $f_4(a,b,c,\phi,\psi)=g(0,0,0,\phi,\psi)=16/3$. Similarly $\lim_{(a,b)\rightarrow (0,0)}$ $f_4(a,b,c,\phi,\psi)=$ $4e^{-2c^2}(e^{2c^2}-1-2c^2-2c^4) /(1-e^{-c^2})^3>0$. Finally we need to check that $\lim_{a\rightarrow 0}f_4(a,b,c,\phi,\psi)>0$. This is considerably more difficult and will be checked over several stages. We have that \begin{equation}\label{expr} \lim_{a\rightarrow 0}f_4(a,b,c,\phi,\psi)=2h(b,c,\theta)/(1-e^{-b^2})^2 (1-e^{-c^2})^2 (1-e^{-w^2}) \end{equation} where \begin{eqnarray}\label{h} h(b,c,\theta)=&&1-(1+2b^2)e^{-2b^2}-(1+2c^2)e^{-2c^2}-e^{-2w^2}\nonumber\\ && +2w^2e^{-2(b^2+c^2)}+2e^{-b^2+c^2+w^2}\cos (2bc\sin\theta)\nonumber\\ && + 4bce^{-(b^2+c^2+w^2)}(\cos\theta \cos (2bc\sin\theta )+\sin\theta \sin (2bc\sin\theta )) \end{eqnarray} and $\theta =\phi+\psi$. Now differentiating $h$ with respect to $\theta$ gives \begin{equation}\label{d_theta} {dh\over d\theta}= 8bce^{-(b^2+c^2+w^2)}S( b c ,\theta) \end{equation} where $S(x,\theta)=\sin\theta\left( \cosh (2x\cos\theta )-\cos (2x\sin\theta )\right) -x\sin (2x\sin\theta )$. We can use $\cosh t\ge 1+t^2/2!+t^4/4!$ for all $t$ and $\sin t\le\sum_{n=0}^4 (-1)^nt^{2n+1}/(2n+1)!$, $\cos t\le \sum_{n=0}^4 (-1)^n t^{2n}/(2n)!$ for $t < 10$ to write \begin{equation} S(x,\theta)\ge {2\over 45}x^4 j(x,\theta)\sin\theta \end{equation} where $j(x,\theta)=15+2x^2\sin^6\theta -6x^2\sin^4\theta+4/7x^4\sin^6\theta-1/7x^4\sin^8\theta -2/63x^6\sin^8\theta$. Note that $j(x,\theta)$ is symmetric about $\pi/2$. Differentiating $j(x,\theta)$ with respect to $\theta$ yields $4/7x^2\sin^3\theta\cos\theta (21\sin^2\theta -42+6x^2\sin^2\theta-2x^2\sin^4\theta-4/9x^4\sin^4\theta )$. The first factor is seen to be zero only at $\theta\in\{ 0,\pi/2 ,\pi\}$. The second factor is quadratic in $x^2$ and has no real roots as $(3-\sin^2\theta)^2+28/3(\sin^2\theta-2)<0$ for all $\theta$. Hence $j(x,\theta)$ takes stationary values for $\theta\in\{ 0,\pi/2 ,\pi\}$. Now $j(x,0)=j(x,\pi)=15$, while $$ j(x,\pi/2)=15-4x^2+{3\over 7}x^4-{2\over 63}x^6 $$ is a cubic in $x^2$ which is easily shown to be decreasing with $x$. $j(2.4,\pi/2)>0$ and thus for $x<2.4$, ${dh\over d\theta}\ge 0$ and $h$ is increasing with $\theta$. We now need to find an increasing lower bound for $h$ when $x\ge 2.4$. For $\theta\in [0,\pi/2]$, so that $\cos\theta\ge 0$, we can get a lower bound for $h$ by using $\cos (2bc\sin\theta )\ge 1-2b^2c^2\sin^2\theta$ and $\sin (2bc\sin\theta )\ge -2bc\sin\theta$. Let this lower bound be $k_1$. Note that $h$ and $k_1$ coincide at $\theta =0$. We have that \begin{equation}\label{k_1} {dk_1\over d\theta}=16bc\sin\theta e^{-(b^2+c^2+w^2)}S_1(b c,b c \cos\theta) \end{equation} where $S_1(x,y)=\sinh^2 y -2y+(2+y)(x^2-y^2)-y^2$. For $x\ge y$ we have that $S_1(x,y)\ge\sinh^2 y-2y-y^2>0$ for $y>1.65$. On the other hand $S_1(x,y)>-2y+2(x^2-y^2)>0$ if $y<(-1+\sqrt{1+4x^2})/2$. Combining these two results we have $S_1(x,y)\ge 0$ for all $y\le x$ if $x>2.1$. For $\theta\in [\pi/2,\pi]$ we have that $\cos\theta\le 0$. We use the same bounds in $h$ as before except we bound $\cos (2bc\sin\theta )$ by $1$ in the term containing $\cos\theta$. Let this bound be $k_2$. Note that $k_1$ and $k_2$ coincide at $\theta =\pi/2$. It is simple to see that \begin{equation}\label{k_2} {dk_2\over d\theta}=8bc\sin\theta e^{-(b^2+c^2+w^2)}S_2(b,c,\theta) \end{equation} where $S_2(b,c,\theta)=2\sinh^2(bc\cos\theta)-4bc\cos\theta+3b^2c^2\sin^2\theta\ge 0$. Hence $k_2$ is increasing with $\theta$. We have shown, when $x\ge 2.4$, that for $\theta\in [0,\pi/2]$, $k_1\le h$ is increasing, and for $\theta\in [\pi/2,\pi]$, $k_2\le h$ is increasing. We have also noted that $h=k_1$ at $\theta =0$ and that $k_1=k_2$ at $\theta =\pi/2$. We have seen that for $x<2.4$, $h$ is increasing. Therefore it remains to check that $h$ is strictly positive at $\theta =0$. Making the change of variables $s=2b^2$ and $r=2b(c-b)$ we have that $$ h(\sqrt{s/2},(r/s+1)\sqrt{s/2},0)={s\over r^2}e^s e^{r^2/s}\ {\tilde h}(s,r) $$ where \begin{eqnarray*} {\tilde h}(s,r)&=&{(e^{r^2/s} - 1)\,s\,(e^{s} - 1 - s)\over r^2} + 2 {e^{ - r}\,s\,(1 - e^{- r})\over r}\\ &&\mbox{\hspace{5.5truecm}} - e^{ - 2r}\,(1 - e^{- s}) - {s\,(1 + s)\,(1 - e^{- r}) ^2\over r^2}. \end{eqnarray*} Let $k_3$ be the lower bound obtained by replacing the first term in ${\tilde h}$ by $(1+r^2/2s)(e^s-1-s)$. We differentiate $k_3$ with respect to $r$ and write it as a power series expansion in $r$; \begin{eqnarray*} s\,r^3\,e^{2r}\,{dk_3\over d r} &=& 2s(1-e^{-s}-s+{s^2\over 2})r^3+ (e^s-1-s-{s^2\over 2}+{5s^3\over 6})r^4\\ & & \mbox{\hspace{40pt}} +2s^2\sum_{n=5}^{\infty}{r^n\over n!}\left(s(2^n-n-2)+2^n-n^2-n-2)\right)\\ & & \mbox{\hspace{60pt}} +(e^s-1-s)\sum_{n=5}^{\infty}{{2^{n-4}r^n}\over {(n-4)!}}. \end{eqnarray*} By using the bound $1-e^{-s}\le s$ it is easy to see that the first term is increasing and thus positive. $e^s-1-s-{s^2\over 2}+{5s^3\over 6}\geq 0$ and $s(2^n-n-2)+2^n-n^2-n-2\geq 0$ for $n\geq 5$. Thus $k_3$ is increasing with $r$. Finally we need to check that the $\lim_{r\rightarrow 0}k_3(s,r)> 0$. $$ \lim_{r\rightarrow 0}k_3(s,r)=2(\cosh s-1-{s^2\over 2})>0 $$ for $s>0$ and therefore $h(b,c,0)>0$ for all non-zero $b$, $c$ and the Lemma is proved.\par\noindent\rightline{$\square$} Recalling (\ref{opb:M_i}) we can write, $$ {\tilde M}^{\lambda}_{\Lambda} =\sum_i M_i\qquad {\hbox { and }}\qquad ({\tilde M}^{\lambda}_{\Lambda})^{-1} =\sum_i (M_i)^{-1}. $$ Let ${\displaystyle {\cal E}= \sup_i\Vert M_i^{- 1}\Vert }$, then $\Vert ({\tilde M}^{\lambda}_{\Lambda})^{-1}\Vert\le {\cal E}$. {\bf Lemma 4.8:} {\sl Let $\delta_\kappa ={1\over 2} e^{-\kappa /64}$. If $\theta>39$ and $q<\theta /13 -3$ then there exists $Q_4$ such that for all $L>Q_4$, $\kappa >64\ln (2L^\theta )$, for all $E\in (-\delta_\kappa ,\delta_\kappa )$, all $\lambda$ with $\vert\lambda\vert <e^{-L^\beta/2}$, $$ \hbox{\BB P}\left( d(E,\sigma (M_{\Lambda_L}^{\lambda}))<e^{-L^\beta}\right) <{1\over L^q}. $$ } {\bf Proof:} If $d_i =\dim {\cal P}_i$ then we have \begin{equation} \Vert M_i^{-1}\Vert\le{{c_{d_i}}\over {\vert\det M _i\vert }}\Vert M_i\Vert^{d_i-1} \end{equation} where $c_{d_i}$ is a constant. Obviously $d_i\in \{ 1,2,3,4 \}$, as a maximal cluster contains 4 impurities. Now, from the previous lemma we have a lower bound for $\vert\det M_i\vert$ for the case where $\lambda =0$. Using the bound $1-\exp (-\kappa x^2)\ge x^2 \exp (-x^2)$ for $\kappa>1$, we can write when $\lambda =0$, \begin{equation} \vert\det M_i\vert\ge\quad C\prod_{m<\, n :\, \zeta_m ,\,\zeta_n\in\, {\cal C}_i}\vert\zeta_m-\zeta_n\vert^2 e^{-\vert\zeta_m-\zeta_n\vert^2}. \end{equation} Let $u=\theta/13$. Then if $\vert\zeta_n-\zeta_{n'}\vert>2/L^u$ for $\zeta_n,\zeta_{n'}\in {\cal C}_i$ with $\zeta_n\ne\zeta_{n'}$, $$ \vert\det M_i\vert\ge C\({4e^{-(3/8)^2}\over L^{2u}}\)^6=C'L^{-12u}. $$ Also, if $\lambda =0$ then $$ \Vert M_i\Vert^{d_i-1}\le\left( \sum_{\{n,m:\zeta_n,\zeta_m\in {\cal C}_i\} } \vert\langle m\vert M_i\vert n\rangle \vert \right)^{d_i-1} <A, $$ for some constant $A$, independent of $\zeta$. So for $\lambda=0$, $\Vert M_i^{-1}\Vert\le C''L^{12u}$. Therefore if $D$ is the diagonal matrix made up of the elements $\lambda\over\omega_n$ with $\vert {\lambda\over\omega_n}\vert<e^{-L^\beta/4}$ for $\{n:\zeta_n\in{\cal C}_i\}$ then for $L$ sufficiently large, by the resolvent identity \begin{equation} {\cal E}=\sup_i\Vert M_i^{-1}\Vert\le \sup_i {\Vert M_i^{-1}\vert_{\lambda =0}\Vert\over 1-\Vert D\Vert\,\Vert M_i^{-1}\vert_{\lambda =0}\Vert}\le L^\theta. \end{equation} The probability for this to occur is greater than $\hbox{\BB P} (\vert\omega_n\vert > e^{-L^{\beta}/4} {\hbox { and }} \vert\zeta_n -\zeta_{n'}\vert>{2\over L^u}$ for all $n,n' {\hbox { such that }} \zeta_n ,\zeta_{n'}\in{\cal C}_i$ with $\zeta_n\ne\zeta_{n'})$ which is greater than $(1-2\rho_be^{-L^{\beta}/4})^4(1-L^{3-u}) >1-L^{-q}$ for $L$ sufficiently large. Let $\delta M^{\lambda}_{\Lambda}=M^{\lambda}_{\Lambda} -{\tilde M}^{\lambda}_{\Lambda}$. Then $$ \langle n\vert \delta M^{\lambda}_{\Lambda}\vert n'\rangle = \cases{ 0 & if $\zeta_n ,\zeta_{n'}$ are in the same cluster, \cr {\pi\over {2\kappa }}P_0(\zeta_n,\zeta_{n'} ) & otherwise. \cr} $$ Since $\kappa\ge\pi /2$ we have, $$ \Vert\delta M^{\lambda}_{\Lambda}\Vert\le e^{-{\kappa/64}}. $$ >From the resolvent identity we get \begin{equation} \Vert ( M^{\lambda}_{\Lambda})^{-1}\Vert\le \Vert ({\tilde M}^{\lambda}_{\Lambda})^{-1}\Vert +\Vert ({\tilde M}^{\lambda}_{\Lambda})^{-1}\Vert\, \Vert \delta M^{\lambda}_{\Lambda} \Vert\, \Vert ( M^{\lambda}_{\Lambda})^{-1}\Vert. \end{equation} If we can make $\Vert ({\tilde M}^{\lambda}_{\Lambda})^{-1}\Vert\, \Vert \delta M^{\lambda}_{\Lambda}\Vert\le {\cal E}e^{-\kappa /64} \le{1\over 2}$ then we get, \begin{equation} \Vert ( M^{\lambda}_{\Lambda})^{-1}\Vert\le 2\Vert ( {\tilde M}^{\lambda}_{\Lambda})^{-1}\Vert. \end{equation} Thus, we have that if $\kappa >64\ln (2L^\theta )$, $$ \Vert ( M^{\lambda}_{\Lambda})^{-1}\Vert\le 2L^\theta, $$ with a probability greater than $1-L^{-q}$ if $L$ is sufficiently large. If $\vert E\vert <{1\over 2} e^{-\kappa /64}$ then $\vert E\vert <1/4L^{\theta}$. So $\Vert (M_{\Lambda}^{\lambda}-E)^{- 1}\Vert <[\Vert (M_{\Lambda}^{\lambda})^{-1}\Vert^{-1}-1/4L^\theta ]^{-1}.$ Hence from above, $$ \hbox{\BB P}\left( \Vert ( M^{\lambda}_{\Lambda}-E)^{-1}\Vert\le 4L^\theta \right) >1-{1\over L^q}. $$ Now, as $E\in\hbox{\BB R}$, $$ d(E,\sigma (M_{\Lambda}^{\lambda}))=\Vert ( M^{\lambda}_{\Lambda}-E)^{-1} \Vert^{-1}, $$ which gives us that $$ \hbox{\BB P}\left( d(E,\sigma (M_{\Lambda}^{\lambda})) \le {1\over {4L^\theta }}\right) <{1\over L^q}. $$ Taking $Q_4$ sufficiently large so that in addition $L^\beta >\ln (4L^\theta)$ we obtain the result.\par\noindent\rightline{$\square$} {\bf Lemma 4.9}: {\sl There exists $Q_5$ such that for all $L>Q_5$, for $q>0$, any $E\in\hbox{\BB R}$, all $\kappa <L^\beta/20$, all $\lambda$ with $\vert\lambda\vert < e^{-L^\beta/2 }$, \begin{equation} \hbox{\BB P} ( d(E, \sigma (M^{\lambda}_{\Lambda_L})) <e^{-L^\beta} ) <{1\over L^q}. \end{equation} } {\bf Proof}: We divide up the points of $\Lambda\cap\hbox{\BB Z} [i]$ into adjacent pairs $\{ n_i,n_i^\prime \}$. Let $Q_i$ be the two dimensional projection onto the space spanned by $\vert n_i\rangle$ and $\vert n_i^\prime \rangle$. Let $$ U_i={1\over \sqrt{2}}\left(\matrix{ 1&-e^{- 2i\kappa\zeta_{n_i}\wedge\zeta_{n_i^\prime}}\cr 1&e^{- 2i\kappa\zeta_{n_i}\wedge\zeta_{n_i^\prime}}\cr}\right)\ ,\quad U_i^* ={1\over \sqrt{2}}\left(\matrix{ 1&1\cr -e^{ 2i\kappa\zeta_{n_i}\wedge\zeta_{n_i^\prime}}&e^{ 2i\kappa\zeta_{n_i}\wedge\zeta_{n_i^\prime}}\cr}\right)\qquad $$ Let $U=\sum_i Q_i U_i Q_i^*$. Then we have, for $n_i, n_i^\prime$ in a pair, \begin{eqnarray*} \langle n_i\vert UM^0 U^*\vert n_i\rangle &=&1+e^{-\kappa\vert\zeta_{n_i}- \zeta_{n_i^\prime}\vert^2}\\ \langle n_i^\prime\vert UM^0 U^*\vert n_i^\prime\rangle &=&1-e^{-\kappa\vert \zeta_{n_i}- \zeta_{n_i^\prime}\vert^2} \end{eqnarray*} Now, if $r=\vert\zeta_{n_i}-\zeta_{n^\prime_i}\vert$, then \begin{eqnarray*} \hbox{\BB P} [r\in (a,b)] &=&\int d^2\zeta_1\int d^2\zeta_2 r(\zeta_1)\, r(\zeta_2)\,1_{\{\vert\zeta_1 -\zeta_2\vert\in (a,b)\} }\\ &\le &\int r(\zeta_1 )d^2(\zeta_1)\int_{\hbox{\BB R}^2} r(\zeta_2)1_{ \{\vert\zeta_1 -\zeta_2\vert\in (a,b)\} }d^2(\zeta_2)\nonumber\\ &\le &2\pi r_b\int_a^b r dr =\pi r_b (b^2 -a^2) =\int_a^b {\tilde \rho}(r)dr \end{eqnarray*} with ${\tilde \rho}(r)=2\pi r_b\, r$. Let $s=e^{-\kappa r^2}$ and $e^{-5\kappa}<a<b<1$. Then $-(\pi r_b\, ds)/(\kappa s)={\tilde \rho}(r)dr$ $$ \hbox{\BB P} [s\in (a,b)]< -{{\pi r_b}\over\kappa}\int_a^b {ds \over s} <{{\pi r_b}\over\kappa} e^{5\kappa}\int_a^b ds={{\pi r_b}\over\kappa} e^{5\kappa}(b-a). $$ The density of $x_{nn}=\langle n\vert UM^0 U^*\vert n\rangle$ is bounded by $\pi r_be^{5\kappa}/\kappa$. So the diagonal terms $1\pm s$ have the same bound. For Borel subsets $B$ of $\hbox{\BB R}$ let $\sigma^\Lambda_n(B) = \langle n \vert E_\Lambda (B) \vert n \rangle$, where $ E_\Lambda (B) $ are the spectral projections of $ UM^0 U^*$, with $\vert\lambda\vert < e^{-L^\beta/2}$. Then by Lemma VIII.1.8 in Ref. 14, $$ \hbox{\BB E}_{x_{nn}} \sigma^\Lambda_n (B) < {{\pi r_be^{5\kappa}}\over\kappa} \int_B dx $$ and therefore $$ \hbox{\BB E} \sigma^\Lambda_n (B) < {{\pi r_be^{5\kappa}}\over\kappa}\int_B dx. $$ As in Proposition VIII.4.11 of Ref. 14, it then follows that for all $E\in\hbox{\BB R}$ and $\epsilon >0$, \begin{equation} \hbox{\BB P} ( d(E, \sigma (M^0_{\Lambda_L})) <\epsilon ) <2{{\pi r_be^{5\kappa}}\over\kappa} \epsilon \vert \Lambda_L \vert <2{{\pi r_be^{5\kappa}}\over\kappa} \epsilon (L+1)^2 <8{{\pi r_be^{5\kappa}}\over\kappa}L^2 \epsilon. \end{equation} if $L\ge 1$. Now, it is easily seen that $\sigma (M^\lambda )\subset \{\, z\, :d(z,\sigma(M^0 ))<\Vert M^\lambda -M^0\Vert\}$. Hence $$ d(E, \sigma (M^\lambda ))\ge d(E, \sigma (M^0)) - \Vert M^\lambda -M^0 \Vert. $$ We can show that $\Vert M^\lambda -M^0 \Vert <e^{-L^\beta /4}$ with a probability $\hbox{\BB P} (\vert\omega_n\vert > e^{-L^\beta/4}\ \ {\hbox { for all }} n\in\Lambda_L\ ) >(1-2\rho_be^{-L^\beta/4})^{(L+1)^2} >1-L^{-2q}$ for all $L>Q_5$ with $Q_5$ sufficiently large. Hence if $d(E, \sigma (M^0_{\Lambda_L}))>\epsilon +e^{-L^\beta/4}$ then $d(E,\sigma (M^\lambda_{\Lambda_L} ))>\epsilon$. So for $\lambda$ with $\vert\lambda\vert < e^{-L^\beta/2 }$, all $E\in\hbox{\BB R}$ and for all $L>Q_5$, \begin{equation} \hbox{\BB P} ( d(E, \sigma (M^{\lambda}_{\Lambda_L})) >\epsilon ) >\left( 1- 8{\pi\over\kappa}r_b L^2 (\epsilon+e^{-L^\beta/4} )e^{5\kappa}\right)\left( 1- {1\over L^{2q}}\right). \end{equation} where we have used $\hbox{\BB P} (A)\ge \hbox{\BB P} (A\vert B)\hbox{\BB P} (B)$. If we have that $L^\beta>20\kappa$ the lemma is proved.\par\noindent\rightline{$\square$} \par Finally, we can bring the two regimes for $\lambda$ together to prove that: {\bf Lemma 4.10}: {\sl There exists $Q_6$ and $\delta_\kappa$ such that for any $q>0$, any $\lambda\in \hbox{\BB R}$, $E\in (-\delta_\kappa ,\delta_\kappa )$, for all $L>Q_6$, $$ \hbox{\BB P}(d(E,\sigma (M^\lambda_{\Lambda_L(0)}))<e^{-L^\beta})<{1\over L^q}. $$ } {\bf Proof}: Choose $\theta>13(q+3)$. Take $Q_6$ larger than $Q_5$ such that $$ 8\rho_b R^2 L^2e^{-L^\beta/2}<{1\over L^q},\quad {\hbox { and }}\quad L^\beta>2^8.5\ln (2L^\theta ) $$ for all $L>Q_6$ and take $\delta_\kappa ={1\over 2} e^{-\kappa /64}$. Let $E\in (-\delta_\kappa ,\delta_\kappa )$ and $\vert\lambda\vert<e^{-L^\beta/2}$. Then by Lemmas 4.8 and 4.9, for all $L>Q_6$, $$ \hbox{\BB P}(d(E,\sigma (M^\lambda_{\Lambda_L(0)}))<e^{-L^\beta})<{1\over L^q}. $$ On the other hand if $\vert\lambda\vert\ge e^{-L^\beta/2}$ then by Lemma 4.5, for all $L>Q_6$ we also have the above inequality.\par\noindent\rightline{$\square$} We finally check that the conditions of Theorem 3.4 are satisfied for $p=3$, $q=25$. By Lemma 4.10 condition (P2) is satisfied for $L>Q_6$ with $\eta =\delta_{\kappa}$ where $\delta_{\kappa}={1\over 2} e^{-\kappa /64}$. Also from Lemma 4.10 $(RA)$ of (P1) is satisfied with probability greater than $1-{1\over L^q}$ for $L>Q_6$. Now in Lemma 4.4 put $u=7$, $\gamma ={1\over 2}$ and let $L_0$ be greater than $Q_3\vee Q_6$ and such that for any fixed $s\in ({1\over 2} ,1)$ (as in the regularity condition) we have $$ {1\over L_0^{25}}+{2\over L_0^4}<{1\over L_0^3},\qquad L_0^{14}(L_0-L_0^s)^{1/2}>64\ln (2L_0^7),\quad {\hbox {\rm and }}\quad L_0^7(L_0-L_0^s)^{1/2}>32L_0. $$ If we choose $\kappa_0=L_0^{4u}/4$ then we have that $$ \vert\langle n\vert \Gamma_{\Lambda_{L_0}}^\lambda (0)\vert n'\rangle\vert\le e^{-\kappa^{1/4}L_0} $$ for all $\kappa >\kappa_0$, all $n$, $n'\in\Lambda_{L_0}$ with a probability greater than $1-2/L_0^4$. Therefore if we take $m_0=\kappa^{1/4}$, $$ \hbox{\BB P}\{\Lambda_{L_0}(0){\hbox { is }}(m_0 ,0)-{\hbox {regular }}\}\ge 1-{1\over L_0^3} $$ where we have used that $\hbox{\BB P} (A\cap B)\ge 1-\hbox{\BB P} (A^c)-\hbox{\BB P} (B^c)$ and condition (P1) is checked. \section{Proof of Theorem 3.2 parts (b) and (c).} In this section we denote by ${\cal L}$ the Lebesgue measure. >From Theorem 3.4, equation 3.5 and an application of Fubini's Theorem we can deduce that with probability one and for ${\cal L}$-a.e. $\lambda$,\ if $\lambda$ is a non-zero generalized eigenvalue of $H(\omega,\zeta)$ then the corresponding eigenfunction decays exponentially. An immediate consequence is that $\sigma_{\hbox {\rm ac }}(H)=\emptyset$. However this does not rule out the existence of singular continuous spectrum. To exclude this we need to show exponential decay for a.e. $\lambda$ with respect to the spectral measure of $H(\omega,\zeta)$. We will use the ideas of Delyon, L\'evy and Souillard$^{17}$. Let $\Lambda\subset\hbox{\BB Z} [i]$ with $\vert\Lambda\vert =N$ as before and define the restriction of $H$ to $\Lambda$ by, $$ H_{\Lambda} =\sum_{n\in\Lambda} \omega_n\vert f_{\zeta_n}\rangle\langle f_{\zeta_n}\vert. $$ If $\psi_k$ are eigenfunctions of $H_{\Lambda}$ with eigenvalues $\lambda_k$, $k=1,\ldots ,N$ and $\Vert \psi_k\Vert =1$ for all $k$, then $$ H_{\Lambda}\psi_k =\lambda_k \psi_k. $$ Define the resolution of the identity of the restriction of $H$ to $H_{\Lambda}$ by $$ E_{\Lambda}(B)=\sum_{k:\lambda_k\in B} \vert \psi_k\rangle\langle\psi_k\vert\ , $$ where $B$ is a Borel subset of $\hbox{\BB R}$ and let $\sigma_{\Lambda}^{\phi ,\phi }=\langle\phi\vert E_{\Lambda}\vert\phi\rangle$ for some $\phi\in {\cal H}_0$. For $\Lambda\nearrow \hbox{\BB Z} [i]$, $\sigma_{\Lambda}^{\phi ,\phi}$ converges weakly as a measure to $\sigma^{\phi ,\phi}$, the spectral measure of $H$. We can write $\sigma_{\Lambda}^{\phi ,\phi}$ explicitly: \begin{equation}\label{opb:specmeas} \sigma_{\Lambda}^{\phi ,\phi} (B)=\sum_{k:\lambda_k\in B}\vert\langle\phi\vert\psi_k\rangle\vert^2\,. \end{equation} As in Section III the eigenvalues $\lambda_k$ of $H_\Lambda$ with eigenfunction $\psi_k$ must satisfy $M_\Lambda^\lambda\xi_k =0$ where $\xi_k (n)=\sqrt {2\kappa /\pi}\ \omega_n\langle f_{\zeta_n}\vert \psi_k\rangle$. Thus we can expect $N$ solutions of $\lambda$ for $\det M_\Lambda^{\lambda }=0$. We will look at solutions as a function of one of the $\omega_n$ only. Using ${{\partial H_{\Lambda} }\over {\partial\omega_n }}=\vert f_{\zeta_n} \rangle\langle f_{\zeta_n} \vert$, a calculation of $\langle\psi_k\vert d/d\omega_n (H_{\Lambda}\psi_k) \rangle$ yields that \begin{equation}\label{opb:dlambdadomega} {{d\lambda_k }\over {d\omega_n }}=\vert\langle f_{\zeta_n} \vert \psi_k\rangle\vert^2\,. \end{equation} Note that if $\langle f_{\zeta_n}\vert\psi_k\rangle=0$ then $\psi_k$ remains an eigenvector for $\lambda_k$ as $\omega_n$ varies and does not contribute to $\sigma_{\Lambda}^{f_{\zeta_n} ,f_{\zeta_n}}$. Also, if $\lambda_k$ is degenerate, we can choose the corresponding orthonormal set of eigenvectors so that only one satisfies $\langle f_{\zeta_n}\vert\psi_k\rangle\ne 0$. From (\ref{opb:dlambdadomega}) we see that each $\lambda_k$ is a monotonic increasing function of $\omega_n$ and from (3.4) that we get $N-1$ solutions of $\lambda_k$ which are identical as $\omega_n\to\pm\infty$. The $N$-th solution corresponds to the $\psi_k$ which tends to $f_{\zeta_n}$ and this value increases from $\lambda =-\infty$ at $\omega_n=-\infty$ to the lowest $\lambda_k$ at $\omega_n=+\infty$ (respectively increases from the highest $\lambda_k$ at $\omega_n=-\infty$ to $\lambda =+\infty$ at $\omega_n=+\infty$) as can be seen from the following argument. We would like to know the behaviour when $\omega_n$ and $\lambda$ both become large. If we expand the determinant we get, $$ \(1-{\lambda\over {\omega_n}}\)\({{(-1)^{N-1}}\over\Pi}\lambda^{N-1} +P(\lambda)\)+\sum_{m\ne n}^N \omega_m {{e^{-2\kappa\vert\zeta_m -\zeta_n\vert^2} (-1)^{N-2}}\over \Pi}\lambda^{N-2}+Q(\lambda)=0 $$ where $\Pi =\prod_{m\ne n}^N\omega_m$, $P(\lambda)$ is a polynomial in $\lambda$ of degree $N-2$ and $Q(\lambda)$ is a polynomial in $\lambda$ of degree $N-3$. We thus get an expression for $\omega_n$ in terms of $\lambda$. $$ \omega_n ={{\lambda^N}\over {\lambda^{N-1}-\sum_{m\ne n}^N \omega_me^{-2\kappa\vert\zeta_m -\zeta_n\vert^2}\lambda^{N-2} +R(\lambda )}} $$ where $R(\lambda )$ is a polynomial in $\lambda^{N-3}$. Thus for $\lambda$ large, $\omega_n \sim\lambda$. This is what we expect if $\psi_k\to f_{\zeta_n}$ as then we have that $d\lambda_k /d\omega_n \to 1$. Now recalling the fact that $\lambda $ is a monotonic increasing function of $\omega_n$ we see that only one among the eigenvalues $\lambda_k$ crosses any $\lambda$, i.e. the range of $\lambda (\omega_n )$ is divided into disjoint open subsets $O_k$ such that $\bigcup_k O_k =\hbox{\BB R}$ and each $\lambda_k$ corresponds to only one of the $O_k$. Therefore there is only one term corresponding to such $\lambda_k$ in the sum (\ref{opb:specmeas}) for $\sigma_{\Lambda}^{f_{\zeta_n} ,f_{\zeta_n}}$. The above results along with (\ref{opb:specmeas}) and (\ref{opb:dlambdadomega}) allow us to make the following change of variables: \begin{eqnarray} \int_{\hbox{\BB R}\times B} \rho(\omega_n )d\omega_n\sigma_\Lambda^{f_{\zeta_n} ,f_{\zeta_n}}(d\lambda)&=&\int_{\hbox{\BB R}}\sum_{k:\lambda_k\in B}\rho (\omega_n)d\omega_n\vert\langle f_{\zeta_n}\vert\psi_k\rangle\vert^2\nonumber\\ &\le &\rho_b\ \#\{\lambda_k\in B\}=\rho_b\,\vert B\vert . \end{eqnarray} Using the weak convergence of $\sigma_\Lambda^{f_{\zeta_n},f_{\zeta_n}}$ to $\sigma_{\phantom \Lambda}^{f_{\zeta_n},f_{\zeta_n}}$ we can therefore write \begin{equation}\label{opb:boundavspecmeas} \int_{\hbox{\BB R}}\rho(\omega_n )d\omega_n\sigma^{f_{\zeta_n} ,f_{\zeta_n}}(B )\le\rho_b\ \vert B\vert , \end{equation} and hence the $\omega_n$-averaged spectral measure $\hbox{\BB E}_{\omega_n}\(\sigma^{f_{\zeta_n} ,f_{\zeta_n}}(d\lambda )\)$ is absolutely continuous with respect to Lebesgue measure. We now use Kotani's ``trick'' (see Ref. 12). In the following, ${\cal B}$ will represent the Borel $\sigma$-field. We will need the following lemma who's proof is elementary. {\bf Lemma 5.1 :}\ {\sl Let $\{ f_n \}$ be a total countable subset of normalised vectors of a Hilbert space ${\cal H}$ and $H$ a self-adjoint operator on ${\cal H}$ with spectral projections $E(\,\cdot\, )$. Let $c_n > 0$, $\sum_n c_n <\infty$ and $\nu =\sum_n c_n \sigma^{f_n,f_n}$, where $\sigma^{f_n,f_n}(\cdot) =\langle f_n\vert E(\cdot )\vert f_n\rangle$. Then for any $B\in {\cal B}$, $\nu (B)=0$ implies that $E(B)=0$.} Let $(\Omega, {\cal F}, \hbox{\BB P} )$ be the probability space corresponding to the $\omega$ and $\zeta$ and let ${\cal F}_n^*$ be the sub $\sigma$-field of ${\cal F}$ generated by all of these variables except $\omega_n$ for some $n\in\hbox{\BB Z} [i]$. If $F(\omega ,\zeta ,\lambda )$ is a nonnegative ${\cal F}^*_n\otimes {\cal B}$ measurable function, then from Proposition VIII.1.4 in Ref. 14 we have that, \begin{equation}\label{opb:kotani} \(\hbox{\BB E}\{\int F(\,\cdot\, ,\,\cdot\, ,\lambda )\,d\lambda\} =0\)\Rightarrow\(\int F(\omega ,\zeta ,\lambda )\sigma^{f_{\zeta_n},f_{\zeta_n}}(d\lambda )=0\ \hbox{\BB P} {\hbox {\rm -a.e. }}\). \end{equation} {\bf Lemma 5.2 :}\ {\sl For $B\in {\cal B}$ let $B\mapsto E (B)$ be the spectral measure of $H$ and let $A\in \cap_{n\in\hbox{\BB Z} [i]}({\cal F}^*_n\otimes {\cal B})$, then, if for a.e. $\lambda\in \hbox{\BB R}$ with respect to Lebesgue measure $\hbox{\BB E}\{ 1_A (\,\cdot ,\,\cdot ,\lambda )\} =0$, then $\hbox{\BB E}\{ E(\{\lambda :\, (\,\cdot ,\,\cdot ,\lambda )\in A\} )\} =0$. } {\bf Proof :}\ Let $A\in {\cal F}^*_n\otimes {\cal B}$. If for a.e. $\lambda$ with respect to Lebesgue measure $\hbox{\BB E}\{ 1_A\{(\,\cdot ,\,\cdot ,\lambda )\} =0$, then by Fubini's Theorem $\hbox{\BB E}\{\int d\lambda 1_A (\,\cdot ,\,\cdot ,\lambda )\} =0$. Combining this with (\ref{opb:boundavspecmeas}) we have that $\hbox{\BB E}\{\int 1_A (\,\cdot ,\,\cdot ,\lambda ) \hbox{\BB E}_{\omega_n} (\sigma^{f_{\zeta_n},f_{\zeta_n}}(d\lambda))\} =0$. We now use the fact that $A\in {\cal F}^*_n\otimes {\cal B}$ with (\ref{opb:kotani}) to move the expectation over $\omega_n$ outside the integral and we obtain \begin{equation} \hbox{\BB E}\{ 1_A\{(\,\cdot ,\,\cdot ,\lambda )\} =0\quad\Rightarrow\quad\hbox{\BB E}\{\int 1_A (\,\cdot ,\,\cdot ,\lambda )\sigma^{f_{\zeta_n},f_{\zeta_n}}(d\lambda)\} =0. \end{equation} Finally, by taking $A\in \cap_{n\in\hbox{\BB Z} [i]}({\cal F}^*_n\otimes {\cal B})$ and $\nu =\sum_n c_n \sigma^{f_{\zeta_n},f_{\zeta_n}}$, where each $c_n > 0$, $\ \sum_n c_n <\infty$ we have that $\hbox{\BB E}\{\int_{\hbox{\BB R}} 1_A (\,\cdot ,\,\cdot ,\lambda )\nu (d\lambda )\}=0$ and from Lemma 5.1 we have the result. Now we have seen at the beginning of this section that if $W$ is the set in $\Omega\times\hbox{\BB R}$ defined by: \begin{eqnarray*} W=&&\{ (\omega ,\zeta ,\lambda ):\ {\hbox {\rm the generalized eigenfunctions of }} H(\omega ,\zeta)\ \\ &&\hskip 6truecm{\hbox {\rm with eigenvalue }}\,\lambda {\hbox { decay exponentially }}\}, \end{eqnarray*} then Fubini's Theorem implies that $W$ is of $\hbox{\BB P}\otimes {\cal L}$ full measure. By taking $W^c$ as $A$ in Lemma 5.2 we have shown that with probability one and $\lambda$-a.e. with respect to the spectral measure, if $\lambda$ is a generalized eigenvalue of $H$ then the corresponding eigenfunctions decay exponentially and hence Theorem 3.2 parts (b) and (c) are proven. \section*{Acknowledgements} We would like to thank T. C. Dorlas for providing a direct proof of Lemma 4.7 in the case of three impurities and for elucidating some aspects of Kotani's trick. One of us (M.S.) acknowledges support from the Forbairt Scientific Research Scheme (Basic Research Scheme SC/1997/621). \newpage \section*{References} \noindent $^1$ J. Desbois, S. Ouvry, C. Texier, Nucl. Phys. B {\bf 500}, 486 (1997). \noindent $^2$ T. C. Dorlas, N. Macris, J. V. Pul\'e, Helv. Phys. Acta {\bf 68}, 330 (1995); J. Math. Phys. {\bf 37}, 1574 (1996). \noindent $^3$ J. M. Combes, P. D. Hislop, Commun. Math. Phys. {\bf 177}, 603 (1996). \noindent $^4$ W-M. Wang, J. Funct. Anal. {\bf 146}, 1 (1997). \noindent $^5$ H. von Dreifus, A. Klein, Commun. Math. Phys. {\bf 124}, 285 (1989). \noindent $^6$ J. Fr\"ohlich, T. Spencer, Commun. Math. Phys. {\bf 88}, 151 (1983). \noindent $^7$ J. Bellisard, A. van Elst, H. Schulz-Baldes, J. Math. Phys. {\bf 35}, 5373 (1994). \noindent $^8$ T. C. Dorlas, N. Macris, J. V. Pul\'e, J. Stat. Phys. {\bf 87}, 847 (1997). \noindent $^9$ T. C. Dorlas, N. Macris, J. V. Pul\'e, Commun. Math. Phys. {\bf 204}, 367 (1999). \noindent $^{10}$ M. Aizenman, S. Molchanov, Commun. Math. Phys. {\bf 157}, 245 (1993). \noindent $^{11}$ M. Scrowston, Ph.D. thesis, National University of Ireland, Dublin (1999). \noindent $^{12}$ S. Kotani: in {\sl Proceedings of the 1984 AMS conference on Random Matrices and their Applications}, Contemp. Math. {\bf 50}, Providence RI (1986). \noindent $^{13}$ J. V. Pul\'e, M. Scrowston, J. Math. Phys. {\bf 38}, 6304 (1997). \noindent $^{14}$ R. Carmona, J. Lacroix: {\sl Spectral Theory of Random Schr\"odinger Operators}. Birkh\"auser - Boston (1990). \noindent $^{15}$ R. Ph. Boas: {\sl Entire Functions}. Academic Press - New York (1954). \noindent $^{16}$ T. C. Dorlas: Private communication. \noindent $^{17}$ F. Delyon, H. L\'evy, B. Souillard, Commun. Math. Phys. {\bf 100}, 463 (1985). \end{document}
\section{Introduction} The properties of 4U\thinspace0142+61\ (White et al. 1987) remained puzzling for a long time, owing to confusion with a nearby pulsating and transient Be/neutron star system RX\,J0146.9+6121 (Motch et al. 1991; Mereghetti {\rc et al. } 1993). The 1--10~keV spectrum is extremely soft (power law photon index of $\sim 4$, White et al. 1987) and led to the initial classification of 4U\thinspace0142+61\ as a possible black hole candidate. ASCA observations provide evidence for a $\sim 0.4$~keV blackbody component contributing $\sim 40$\% of the 0.5--10 keV band X--ray flux (White et al. 1996). The X--ray luminosity of 4U~0142+61 has not shown substantial secular variations around an average value of $\sim~6 \times 10^{34}$~erg~s$^{-1}$ (assuming a distance of 1~kpc). Despite the small error box (5\arcsec\ radius), no optical or IR counterpart has yet been identified, down to $V<24$, $R<22.5$, $J<20$ and $K<17$ (Steinle et al. 1987; White et al. 1987; Coe \& Pightling 1998). These limits rule out the presence of a massive companion. Using data from the EXOSAT archive, Israel et al. (1994) discovered pulsations at 8.7 s, which were later confirmed with ROSAT (Hellier 1994). No delays in the pulse arrival times caused by orbital motion were found, with upper limits on $a_{\rm x}$\,sin $i$ of about $\sim 0.37$~lt--s for orbital periods $P_{\rm orb}$ between 7~min and 12~hr (Israel et al. 1994). {\rc Tighter} upper limits on the $a_{\rm x}\sin i$ ($\sim 0.26$~lt--s for 70\,s\,$\leq P_{\rm orb} \leq$ 2.5\,days) have been recently obtained with a RXTE\ observation (Wilson et al. 1998). This yielded strong constraints on the orbital inclination and the mass of the possible companion star in the case of normal or helium main sequence star and giants with helium core. A white dwarf companion would be compatible both with current optical photometric and pulse arrival time limits. \begin{table*} \caption[]{4U\thinspace0142+61\ \BSAX\ Observation log} \begin{tabular}{cccccccc} \hline Start Time & Stop Time & MECS T$_{\rm exp}$ & Active MECSs & Count Rate LECS & MECS & Off--axis & Obs \\ & & s & \# & c/s & c/s & \arcmin\ & \\ \hline 97 Jan 03~~05:47 & Jan 04~~02:09 & 48226 &3& 1.03$\pm$0.01& 1.70$\pm$0.01 &5 & A\\ 97 Aug 09~~23:13 & Aug 10~~07:38 & 16757 &2& 1.01$\pm$0.02 & 1.47$\pm$0.02&3 & B\\ 98 Jan 26~~12:13 & Jan 27~~00:23 & 21785 &2& (4.2$\pm$0.4)$\times$10$^{-2}$&0.47$\pm$0.01 &20 & C\\ 98 Feb 03~~09:11 & Feb 04~~02:37 & 31150 &2& 0.97$\pm$0.02& 1.36$\pm$0.01&5 & D\\ \hline \hline \end{tabular} \\ ~ \\ Note --- Count rates are not vignetting corrected. Vignetting is a factor 1.10, 1.20 and 3.0 for an off--axis angle of 3, 5 and 20 arcmin, rescpectively. \end{table*} \begin{figure*}[tbh] \centerline{ \psfig{figure=specA.ps,width=6cm,height=8cm} \psfig{figure=specD.ps,width=6cm,height=8cm}} \caption{LECS and MECS energy spectra of 4U\thinspace0142+61\ during observation A (left panel) and D (right panel). The best fit spectral model (PL+BB; see text for details) is also shown together with the \chisq\ residuals.} \end{figure*} The EXOSAT and ROSAT period measurements, obtained in 1984 and 1993, provide a spin--down rate of $\sim 2.1\times10^{-12}$~s~s$^{-1}$. The properties of 4U\thinspace0142+61\ are similar to those of a small group of ``anomalous'' X--ray pulsars (AXPs), with spin periods within a narrow range (6--12\,s; Mereghetti \& Stella 1995). Among these are 1E\,2259+586, 1E\,1048.1--5937, 1E\,1841--045 (Vasisht \& Gotthelf 1997), 1RXS\,170849--400910 (Sugizaki et al. 1997) and AX\,J1845--045 (Torii et al. 1998; Gotthelf \& Vasisht 1998). We present here a detailed analysis of the \BSAX\ Narrow Field Instruments (NFIs) observations of the AXP 4U\thinspace0142+61. We confirm the presence of a blackbody spectral component in the soft X--ray spectrum as seen in the ASCA data (White et al. 1996) and in other two ``anomalous'' X--ray pulsars: 1E\,2259+589 (Corbet et al. 1995; Parmar et al. 1998) and 1E\,1048.1--5937 (Oosterbroek et al. 1998). We also present the results of pulse phase spectroscopy and the pulse period history of 4U\thinspace0142+61. {\rc We also discuss the timing analysis from a serendipitous Rossi X--ray Timing Explorer (RXTE) observation of 4U\thinspace0142+61.}\\ \section{\BSAX\ Observations} Results from the Low--Energy Concentrator Spectrometer (LECS; 0.1--10~keV; Parmar et al. 1997) and Medium--Energy Concentrator Spectrometer (MECS; 1.3--10~keV; Boella et al. 1997) on--board \BSAX\ are presented. {\rc The MECS consists of three identical grazing incidence telescopes with imaging gas scintillation proportional counters in their focal planes. The LECS uses an identical concentrator system as the MECS, but utilizes an ultra-thin (1.25~$\mu$m) detector entrance window and a driftless configuration to extend the low-energy response to 0.1~keV. The fields of view (FOV) of the LECS and MECS are circular with diameters of 37\arcmin\ and 56\arcmin\, respectively. The energy resolution of both instruments is $\simeq$ $8.5\sqrt{(6keV/{\rm E})}$ \% full--width half maximum (FWHM), where E is the energy.} \begin{figure}[tbh] \centerline{\psfig{figure=uspecA.ps,width=6cm,height=8cm}} \caption{4U\thinspace0142+61\ LECS and MECS unfolded spectrum for observation A. The power--law and blackbody components are shown (dotted and stepped lines).} \end{figure} 4U\thinspace0142+61\ was observed by \BSAX\ four times between January 1997 and February 1998 (see Table\,1). One of the goals of the program was to monitor the possible X--ray flux variations on a time scale of months with two 40\,ks pointings. {\rc However} between the first and the subsequent observations one of the three MECS units failed (1997 May 9), and data of observations B, C and D were obtained with the remaining two MECS units. Moreover during the second and third observations \BSAX\ experienced pointing failures resulting in a shorter effective exposure time. \subsection{Spectral analysis} Spectra were obtained centered on the position of 4U\thinspace0142+61\ using an extraction radius of 8\arcmin\ for both the LECS and MECS. Background subtraction was performed using standard blank field exposures. The average background subtracted source count rates are reported in Table\,1 for the four observations, together with \ the off--axis angle and number of working MECS. The PHA spectra were rebinned so as to have $>$40 counts in each energy bin and reliably adopt a minimum $\chi^2$ technique for model fitting. All the bins which were consistent with zero after background subtraction were rejected. Moreover the MECS spectra were restricted to the 1.8--10~keV range. Data from observation B and C were not used for spectral analysis purposes owing to poor statistics. A constant factor free to vary within a predetermined range was applied in the fitting to allow for known normalization differences the LECS and MECS. In order to compare our results with previous observations, the spectra were first fit with two models: (i) an absorbed power--law, and (ii) an absorbed power--law plus blackbody (see Table\,2). The power--law model gave an unsatisfactory description of the spectra with \rchisq\ = $\chi^2$/degrees of freedom (hereafter dof) of 1.6 (347 dof) and 1.9 (396) for observations A and D, respectively. We note that among simple single component spectral models, the power--law gives the lowest reduced $\chi^2$; this is for a photon index $\Gamma$ = 4.37$\pm$0.03 (A) and 4.55$\pm$0.03 (D; 90\% confidence level uncertainties are used throughout the paper). The power--law plus blackbody model gave \rchisq\ = 1.28 (344) and 1.07 (290) for observations A and D, respectively (see Table\,2 for details). An F--test shows that the inclusion of the blackbody component is highly significant (probability of $\sim$10$^{-26}$ and $\sim$10$^{-32}$ for obs. A and D, respectively). The best fit two--component spectra are plotted in Fig.\,1. In Fig.\,2 the unfolded energy spectrum for observation A is shown together with the contributions of the two spectral components, the power--law and the blackbody. \begin{table*} \begin{center} \caption[]{\BSAX\ phase averaged fit of 4U\thinspace0142+61.} \begin{tabular}{lcccccc} \hline Spectral Parameter & Obs. A & Obs. A$^b$ & Obs. D & Obs. D$^b$ & Obs. A$^a$+D & Obs. (A+D)$^b$\\ \hline N${\rm _H}$ (10$^{22}$\hbox {\ifmmode $ atoms cm$^{-2}\else atoms cm$^{-2}$\fi})\,............................... & 1.11$\pm$0.07 & 1.1 $\pm$0.1 & 0.98$\pm$0.06 & 1.2$\pm$0.1 & 1.12$\pm$0.06 & 1.0$\pm$0.1\\ $\Gamma$\,.............................................................. & 3.86$\pm$0.06 & 3.8 $\pm$0.3 & 3.58$\pm$0.12 & 4.0$\pm$0.2& 3.95$\pm$0.05 & 3.6$\pm$0.3\\ PL flux (erg s$^{-1}$cm$^{-2}$; 0.5--10 keV)\,........& 7.4$\times10^{-11}$ & 6.0$\times10^{-11}$ &7.3$\times10^{-11}$ & 8.4$\times10^{-11}$ &7.9$\times10^{-11}$ & 7.7$\times10^{-11}$\\ BB kT (keV)\,\,........................................... & 0.42$\pm$0.02 & 0.37$\pm$0.02& 0.36$\pm$0.01 & 0.41$\pm$0.04& 0.40$\pm$0.01 &0.36$\pm$0.02\\ BB radius (km @ 1\,kpc)..............................& 1.5$\pm$0.2 & 2.2$\pm$0.5 & 2.1$\pm$0.2 & 1.5$\pm$0.5 & 1.8$\pm$0.2& 2.3$\pm$0.4 \\ BB flux (erg s$^{-1}$cm$^{-2}$; 0.5--10 keV)\,........& 2.9$\times10^{-11}$ & 2.6$\times10^{-11}$ & 4.1$\times10^{-11}$ & 2.1$\times10^{-11}$ & 3.0$\times10^{-11}$ & 2.3$\times10^{-11}$\\ $\chi^2$/dof\,...................................................... & 440/344 & 214/209 & 309/290 & 203/187 & 455/415 & 283/279 \\ $L_X$ (10$^{34}$ erg s$^{-1}$ @ 1\,kpc; 0.5--10 keV).......& 7.9 & 8.3 & 6.3 & 7.5 & 9.3 & 5.7 \\ \hline \hline \end{tabular} \end{center} Note --- Fluxes are not corrected for the interstellar absorption. Flux uncertainties are about 10\%. The source luminosities were derived by setting N$_H$ = 0. \\ $^a$ Only MECS2 and MECS3 considered.\\ $^b$ Only LECS data considered. \end{table*} \begin{table*} \caption[]{Period history for 4U\thinspace0142+61. 90\% uncertainties are reported.} \begin{center} \begin{tabular}{llccl} \hline Mission &Instrument & Period & Date & Reference\\ & & (s) & (year) & \\ \hline {\em Einstein}&SSS & 8.68707$\pm$0.00012 & 1979.16 & White et al. 1996 \\ {\em Einstein}&MPC & 8.68736$\pm$0.0007 & 1979.67 & White et al. 1996 \\ {it EXOSAT}&ME & 8.68723$\pm$0.00004 & 1984.66 & Israel et al. 1994 \\ {\em ROSAT}&PSPC & 8.68784$\pm$0.00004 & 1993.12 & Hellier 1994 \\ ASCA&GIS & 8.68791$\pm$0.00015 & 1994.72 & White et al. 1996 \\ RXTE&PCA & 8.6881$\pm$0.0002 & 1996.32 & this work \\ RXTE&PCA & 8.688068$\pm$0.000002 & 1996.24 & Wilson et al. 1998 \\ \BSAX&MECS & 8.68804$\pm$0.00007 & 1997.01 & this work\\ \BSAX&MECS & 8.6882$\pm$0.0002 & 1997.61 & this work\\ \BSAX&MECS & 8.6883$\pm$0.0001 & 1998.10 & this work\\ \hline \hline \end{tabular} \end{center} \end{table*} We also fit the LECS and MECS spectra of observation A with other two--component spectral models, such as a power--law with a cut--off (\chisq/dof = 453.2/344), two blackbodies (\chisq/dof = 503.1/344) and a broken power--law (\chisq/dof = 400.6/344). All {\rc these} models gave substantially worse fits than the power--law plus blackbody model. Similar results were obtained for the spectrum of observation D. Since the blackbody peaks are close to the lower end of the MECS energy range we also checked these results by analysing the LECS data only (which cover a wider spectral range than the MECS;see Table 2 and Discussion). We also checked the stability of the results obtained by rebinning by a factor of $\sim$3 the PHA channels in the LECS and MECS spectra. Again all the bins which were consistent with zero after background subtraction were rejected from the analysis. No significant variation in the spectral parameters and uncertainties were found for either models. \begin{figure}[hbt] \centerline{\psfig{figure=allphase.ps,width=6cm,height=12cm}} \caption{The 4U\thinspace0142+61\ MECS light curves for obs. A (first three panels) folded to the best period (P=8.68804\,s) in the 0.5--1.5 keV, 1.5--4 keV and 4--10 keV energy bands. Zero phase was (arbitrarily) chosen to correspond to the minimum in the 1.5--4 keV folded light curve. The results of the pulse phase spectroscopy of obs. A (last two panels) are also reported for two spectral parameters. For clarity two pulse cycles are shown.} \end{figure} The data from the {\rc High Pressure Gas Scintillation Proportional Counter (HPGSPC) and the Phoswich Detector System (PDS)} did not hold any useful information on 4U\thinspace0142+61. In fact, due to the large FOVs of these intruments and the steep spectrum of 4U\thinspace0142+61, the counts above 10 keV were likely dominated by the nearby source RX J0146.9+6121. \subsection{Pulse timing, folded light curves and phase resolved spectroscopy} The arrival times of the 0.5--10~keV photons from 4U\thinspace0142+61\ were corrected to the barycenter of the solar system and 1\,s binned light curves accumulated for each observation. The average count rates are reported in Table\,1. The MECS counts were used to determine the 4U\thinspace0142+61\ pulse period. The data from observations A, B and D were divided into 6, 4, and 5 time intervals, respectively and for each interval the relative phase of the pulsations was determined. These phases were then fit with a linear function giving a best--fit period of $8.68804 \pm 0.00007$~s for observation A (see Table\,3). For observation B a period of $8.6882 \pm 0.0002$~s was obtained, while during observation C pulsations were not detected owing to poor statistics ($\sim$ 11000 photons) to detect such a weak ($\sim$6\% pulsed fraction) signal. Finally for observations D we determined a period of 8.6883$\pm$0.0001\,s. The \BSAX\ pulse period values are plotted in Fig.\,4 together with the previous measurements (see Sect.\,3 for the RXTE\ measurement). The background subtracted light curves from observation A, folded at the best period in different energy ranges (Fig.\,3; first three panels) show a double--peaked profile (see also White et al. 1996). The pulsed fraction (semiamplitude of modulation divided by the mean source count rate) was 7.1$\pm$1.5\%, 7.5$\pm$0.5\% and 13$\pm$2\% in the 0.5--1.5 keV, 1.5--4.0 keV and 4.0--10 keV energy bands, respectively. Similar results are obtained for observations B and D. \begin{figure}[hbt] \centerline{\psfig{figure=0142his_fit.ps,width=6cm,height=8cm}} \caption{Pulse period history of 4U\thinspace0142+61\ as a function of time. 1$\sigma$ uncertainties have been used.} \end{figure} Flux variations on several timescales are usually displayed by high accretion rate X--ray pulsars; we found no short or long--term variations in any of the \BSAX\ accumulated light curves. The upper two panels of Fig.\,5 show the pulsed fraction versus energy during observation A for the first two harmonics. These values were obtained by fitting the corrisponding light curves with two sinusoidal functions. The first harmonic shows a nearly constant value ($\sim$ 6\%) up to 4\,keV, while at higher energies increases to about 15\%. A constant value of 5--7\% is inferred for the second harmonic. We calculated also the root mean square variability (rms; defined as $\sqrt{V_{obs}-V_{exp}}$ devided by the mean source count rate, where $V_{obs}$ and $V_{exp}$ are the observed and expected variance, respectively) of the folded light curve at different energies (third panel). \begin{figure}[tbh] \centerline{\psfig{figure=pulseAnew.ps,width=6cm,height=11cm}} \caption{Pulsed fraction of the first two harmonics of 4U\thinspace0142+61\ as a function of the energy for observation A (upper two panels), together with the rms and the power--law to total flux ratio (lower two panels). See text for details.} \end{figure} Finally the ratio between the power--law and the total flux for the blackbody plus power--law spectral model is shown (lowest panel). From the comparison of these quantites we can infer that: (i) since the rms variability of the folded light curves is $\pi^{-1/2}$ times the geometric sum of the amplitudes in all available harmonics (mainly the 1$^{st}$ and 2$^{nd}$ in the present case), it is apparent that the behaviour is consistent with the derived amplitudes of the 1$^{st}$ and 2$^{nd}$ harmonics; (ii) there is evidence for an increase of the pulsed fraction ($\sim$ 2$\sigma$ confidence level) at energies above 5 keV; (iii) there is not a simple relationship between the pulsed fraction and the flux in either of the two spectral component adopted in our analysis. A set of four phase--resolved spectra (phase boundaries 0.06, 0.26, 0.5, 0.78) were accumulated for observation A (i.e. the observation with the highest number of source counts). These were then fit with the power--law plus blackbody model described in Sect.\,2.1, with N$_H$ fixed at the phase--averaged best--fit value. Initially, the blackbody temperature was fixed at its phase--averaged best--fit value and only the power--law parameters and blackbody normalisation were allowed to vary. The fits were then repeated with $\Gamma$ fixed and blackbody parameters and power--law normalisation free {\rc (see Fig.\,3). No significant changes were detected for $\Gamma$ and $kT$ as a function of the pulse phase. Similar results were obtained for the fluxes of the two spectral component.} \section{RXTE\ Observation} The 4U\thinspace0142+61\ position was included in a RXTE\ observation pointed at the nearby high--mass X--ray binary pulsar RX J0146.9+6121 (from 1996 March 28 11:16:48 to 22:09:20; $\sim$20~ks of effective exposure time). The results presented here are based on data collected in the so called ``Good Xenon" operating mode with the Proportional Counter Array (PCA, Jahoda et al. 1996). The PCA consists of 5 proportional counters operating in the 2--60 keV range, with a total effective area of approximately 7000~cm$^{2}$ and a field of view, defined by passive collimators, of $\sim1\deg$ FWHM. The data consist of the time of arrival and the pulse height of each count. In order to minimise the contamination from RX\,J0146.9+6121 (a much harder spectrum X--ray pulsar), we considered only photons in the 2--4 keV energy interval. In order to reduce the background, only the counts detected in the first Xenon layer of each counter were used. \begin{figure}[tbh] \centerline{\psfig{figure=xte0142.ps,width=6cm,height=4cm}} \caption{The 4U\thinspace0142+61\ RXTE\ 2--4 keV light curve folded to the best fit period (P=8.6881\,s).} \end{figure} We obtained for 4U\thinspace0142+61\ a best pulse period of 8.6881$\pm$0.0002\,s (see Table\,3 and Fig.\,4). Figure\,6 shows the corresponding folded light curve. Despite the uncertainties in the background subtraction (the contribution from RX\,J0146.9+6121 is difficult to estimate), it is apparent that the RXTE/PCA pulse profile of 4U\thinspace0142+61\ is similar to that observed with \BSAX. The RXTE\ value we obtained is consistent to within the statistical uncertainties with that of Wilson et al. (1998; energy range 3.7--9.2 keV). The somewhat different pulse shape is likely due to the different energy band used. \section{Discussion} Several models have been proposed in order to explain the nature of the ``anomalous'' X--ray pulsars. Mereghetti \& Stella (1995) proposed that these sources form a homogeneous subclass of accreting neutron stars, perhaps members of low mass X--ray binaries (LMXBs), which are characterized by lower luminosities ($10^{35}-10^{36}$~erg~s$^{-1}$) and higher magnetic fields ($B\sim10^{11}$~G) than classical LMXBs. However the lack of evidence for a binary nature from any of these systems (Mereghetti et al. 1998; Wilson et al. 1998) argues in favor of models in which the X--ray emission originates from a compact object that is not in an interacting binary system. Van Paradijs et al. (1995) proposed that AXPs are young population I objects, which originate from the evolution of short orbital period High Mass X--ray Binaries (HMXBs), following the expansion of the massive star and the onset of unstable Roche Lobe overflow, before central hydrogen is exhausted (see also Cannon et al. 1992). The resulting common--envelope phase should {\rc cause the neutron star to spiral--in and disrupt the companion star after the so--called Thorne--Z\.ytkov stage.} Therefore, these sources should consist of an isolated neutron star accreting matter from a residual disk with a mass in the $10^{-3}-1$~M$_{\odot}$ range. The blackbody component in 4U\thinspace0142+61\ and, likewise, other AXPs has been interpreted as evidence for quasi--spherical accretion onto an isolated neutron star formed after common envelope and spiral--in of a massive short--period (P$_{orb}$$\leq$ 1\,yr) X--ray binary. In this case the remains of the envelope of the massive star might produce two different types of matter inflow: a spherical accretion component with low specific angular momentum giving rise to the blackbody component and a high specific angular momentum component which forms an accretion disk and is likely responsible for the power--law emission (Ghosh et al. 1997). Although some problems remain open in this scenario, we note that a post common--envelope evolutionary scenario for 4U0142+614 is also suggested by the inference that about half of the X--ray absorbing material is close to the source (White et al. 1996). In this context it is interesting to note that independent evidence supports the view that the binary X--ray pulsars 4U1626--67 and HD49798 were formed following a common envelope and spiral--in evolutionary phase (see Angelini et al. 1995 and Israel et al. 1997 and references therein). The different properties of the companion stars in these two cases may reflect the fact that unstable mass transfer set in at different evolutionary phases in the nuclear evolution of their progenitors (in turn, reflecting different initial masses and orbital periods; see also Ghosh et al. 1997). Thompson \& Duncan (1993, 1996) proposed that the AXPs are ``magnetars'', neutron stars with a superstrong magnetic field ($\sim10^{14-15}$ G). This proposal is supported by the similarity in the pulse periods (8.05\,, 7.5\,s and 5.16\,s) and of \.P values measured in the Soft $\gamma$--ray Repeaters SGR0520--66 , SGR1806-20 and SGR1900+14, respectively (Kouveliotou et al. 1998; Hurley et al. 1998). If this connection proved correct, AXPs, would be quiescent soft $\gamma$--ray repeaters. Heyl \& Hernquist (1997) argued that their emission may be powered by the cooling of the core through a strongly magnetised envelope of matter (made up mainly by hydrogen and helium). Pulsations would originate from a temperature gradient on the surface of the star. Moreover Heyl \& Hernquist (1998) showed that in this scenario, spin--down irregularities, observed in the period history of two ``anomalous'' X--ray pulsars (1E\,2259+58 and 1E\,1048--59), may be simply accounted for with glitches like those observed in young radio pulsars. The 0.5--10 keV spectrum of 4U\thinspace0142+61\ is well modelled by the sum of an absorbed steep power--law and a low--energy blackbody. The latter component was introduced by White et al. (1996) based on the ASCA data; the corresponding blackbody radius and flux were $\sim2.4\pm0.3$\,km (at 1 kpc) and $\sim$40\% of the total, respectively. {\rc Neither the power--law photon index nor the blackbody temperature show} evidence of changes across different \BSAX\ observations. We found marginal evidence (at about 2$\sigma$ confidence level) for changes relative to previous observations. A comparison of the ASCA and \BSAX\ results show that: (i) the power--law photon index increases by about 4\% (observations A and D); (ii) the blackbody radius decreased by about 37\% (A) and 12\% (D); (iii) the blackbody flux decreased to $\sim$30\% (A) and $\sim$35\% (D) of the total flux therefore showing a $\sim$10\%--5\% decrease with respect to that of the ASCA observation; (iv) the 0.5--10 keV total flux decreased of about 15--12\%. By fixing the parameters N$_H$, $kT$ and $\Gamma$ to the values inferred by ASCA, the \BSAX\ observation A spectrum gives a \rchisq/dof of 1.9/348. In this case the blackbody component accounts for $\sim$30\% of the total flux ($\sim$10\% lower than that of ASCA). As an additional test we also merged the data for observation A (MECS 2 and 3 only) and D and fit them with the power--law plus blackbody model (see Table\,2). While a small change in the spectral parameters is found relative to observations A and D, separately, the blackbody flux is $\sim$30\% of the total. To remove possible effects introduced by the vicinity of the lower end of the energy range of the MECS and the peak of the blackbody component, we also fitted the spectrum of observations A, D and A+D by using only the data as accumulated by the LECS, the energy band of which uninterruptely covers the range 0.5--9.0 keV and allows a better fit in the energy interval (1--4 keV) where the power--law and the blackbody components overlap. Again we obtained results similar to the LECS+MECS case, but with a larger uncertainty. All these results point to a marginal variation of the spectral parameters of 4U\thinspace0142+61, if any. The \BSAX\ data {\rc suggest} that the pulsed fraction for energies above 4\,keV decreased (from 25\%$\pm$5\% with ASCA to 13\%$\pm$2\% with \BSAX; 90\% uncertainties). For these energies the counts are almost entirely dominated by the power--law component (in the spectral model assumed). The pulse periods measured by \BSAX\ show that 4U\thinspace0142+61\ has continued its secular spin--down during 1996--1998 (see Fig.\,4). The period derivative inferred from the \BSAX\ observations alone ($\sim$6.0$\pm^7_{5.1}$ $\times$10$^{-12}$ s s$^{-1}$) is consistent with the average spin--down rate ($\sim$2$\times$10$^{-12}$ s s$^{-1}$) inferred over the 19\,yr span of the historical dataset (note that in the period list we did not include the low significance detections inferred on 1985 November 11 and December 11 with EXOSAT ME, and on 1991 February 13 with ROSAT HRI; Israel et al. 1994). The inferred \.P for 4U\thinspace0142+61\ is consistent with the uncertainties of all pulse period measurements except, perhaps, one (the 1979 {\em Einstein}\ SSS one; see Fig.\,4) implying that, so far, no ``glitches'' have been observed yet. \begin{acknowledgements} We would like to thank Giancarlo Cusumano for providing the MECS off--axis matrices and the \BSAX\ Mission Planning for their constant help. The authors also thank K. Long the comment of which helped to improve an earlier version of this paper. This work was partially supported through ASI grants. \end{acknowledgements}
\section*{acknowlegments} This work is supported by Grant from Natural Science Foundation of China.
\section{Introduction} \label{sec:intro} In general relativity it is thought that the assumption of the field equations is sufficient to ensure that gravitational radiation is null (i.e. travels at the speed of light). It is not necessarily the case that gravitational radiation is null \cite{bi:mdr87}, but in general relativity four facts support this view. The {\it first} is that in the linear approximation, with the harmonic gauge, the field equations reduce to $\Box h_{ab}=0$, which has only null wave solutions. The weak field linear approximation may no do justice to the non;inearity of the fielf equations; for example, on a de Sitter background, with cosmological constant $\Lambda$, perturbations obey the Fierz-Pauli equation $(\Box-m^2)h_{ab}=0$, with mass $m^2=2\Lambda/3$ \footnote{footnote added 1999, this calculation is done section 6 gr-qc/9812091}. The {\it second} fact is that exact solutions of Einstein's field equations which are good models of gravitational radiation represent null radiation. The {\it third} is that the charateristics (or shock wave soltuions) of the field equations are null; however, this is not a particularly compelling reason - for example, the characteristics of the massive Klein-gordon equation are null \cite{bi:synge}. The {\it fourth} is that as a consequence of choosing a metric geometry (where $\nabla_cG_{ab}=0$), there is the equation$\Box g_{ab}=0$, which appears to be massless; this has led to the construction of nonmetric massive theories of gravity obeying the equation $(\Box+m)g_{ab}=0$ \cite{bi:mdr86}. One of the most important variants of general relativity is the theory with quadratic terms added to the Lagrangian. Einstein's field equations of general relativity can be derived from Hilbert's Lagrangian \cite{bi:hilbert}. This Lagrangian was generalized to include terms of higher order in the early days of relativity, by Pauli \cite{bi:pauli}, Bach \cite{bi:bach} and others. Recently \cite{bi:schmidt} \cite{bi:mdr91}, the relevance of quadratic Lagrangian theories to the shape of galaxies has been discused. When quadratic Lagrangian theories are linearized \cite{bi:stelle} \cite{bi:BC}, it is found that the field equations reduce to two wave equations with masses fixed by the values of the coupling constants. In one case the massis positive, leading to slower-than-light waves, and in the other it is negative, leading ot faster-than-light waves. The initial value formulation of quadratic Lagrangian theory \cite{bi:noakes}, suggests that these masses are present in the full nonlinear theory. Here the existence of exact nonnull wavelike solutions to the quadratic Lagrangian theory is investigated. The method used is to start with a modification of the Peres' wave \cite{bi:peres}, where the modification entails changing from Peres' dependence on a null coordinate to a nonnull coordinate. Using of coordinate transformations, when the coordinate is timelike the metric is shown to be equivalent to a metric with the lapse $N$ as the only varying component of the metric, and when the metric is spacelike the metric is shown to be static. The Peres metric itself has been discussed in the context of quadratic Lagrangian theories by Buchdahl \cite{bi:buchdahl} and Madsen \cite{bi:madsen}; the metric they use has cylindrical symmetry and depends on a null coordinat, whereas the solutions here do not have these properties. The conventions used are those of Hawking and Ellis \cite{bi:HE}. \section{Quadratic Lagrangians} \label{sec:QL} The action is taken to be of the form \begin{equation} S=\int g^\frac{1}{2}[{\mathcal L}_m+\kappa^{-2}(R-2\Lambda)+bC^2+pR^2]dx^4, \label{eq:2.1} \end{equation}} %\indent where ${\mathcal L}_m$ is the matter Lagrangian, the $R$ term is the Hilbert \cite{bi:hilbert} action, $\Lambda$ is the cosmological constant, the $C^2$ term is the Bach \cite{bi:bach} action with $C^2=C_{abcd}C^{abcd}$ being the square of the Weyl tensor, and the $R^2$ term is the Pauli \cite{bi:pauli} action. Such nomenclature is historically simplistic; for example, Pauli was considering theories with nonmetric connection. Linearization \cite{bi:stelle} \cite{bi:BC} shows that the coupling constants can be identified with the masses \begin{equation} \kappa^{-2}=\frac{cm^2_{PL}}{16\pi\hbar},~~~ m^2_{PL}=\frac{c\hbar}{G},~~~ b=\frac{-1}{2\kappa^2m^2_{RB}},~~~ p=\frac{+1}{6\kappa^2m^2_{WP}}. \label{eq:2.2} \end{equation}} %\indent For $\Lambda=0$, there are eight physical degrees of freedom, two for the massless graviton corresponding to the Hilber term, one massive scalar corresponding to the Pauli term, and five massive spin $2$ poltergeists corresponding to the Bach term. The sign of these terms can also be considered from the point of view of cosmological stability \cite{bi:CT}. Varying \ref{eq:2.1} gives the field equations \begin{equation} \kappa^{-2}(G_{ab}+\Lambda g_{ab})+M_{ab}+X_{ab}=T_{ab}, \label{eq:2.3} \end{equation}} %\indent where \begin{eqnarray} &&G_{ab}=R_{ab}-\frac{1}{2}g_{ab}R,\label{eq:2.4}\\ &&M_{ab}=-2b\Box R_{ab} +2\left(p+\frac{b}{3}\right)R_{;ab} -\left(2p-\frac{b}{3}\right)g_{ab}\Box R,\label{eq:2.5}\\ &&X_{ab}=4bR_{acdb}R_{..}^{cd} -2\left(p-\frac{2b}{3}\right)RR_{ab}\nonumber\\ &&~~~~~~~~~ +g_{ab}\left[bR_{cd}R_{..}^{cd} +\frac{1}{2}\left(p-\frac{2b}{3}\right)R^2\right]. \label{eq:2.6} \end{eqnarray} The trace of \ref{eq:2.3} is \begin{equation} \kappa^{-2}(4\Lambda-R)-6p\Box R=T, \label{eq:2.7} \end{equation}} %\indent with $T=T^{~c}_{c.}$, which leads to the alternative form of the field equations \begin{equation} \kappa^{-2}(R_{ab}-g_{ab}\Lambda)+X_{ab}+Z_{ab}=S_{ab}, \label{eq:2.8} \end{equation}} %\indent where \begin{eqnarray} &&Z_{ab}=-2b\Box R_{ab} +\left(p+\frac{b}{3}\right)[2R_{;ab}+g_{ab}\Box R],\label{eq:2.9}\\ &&S_{ab}=T_{ab}-\frac{1}{2}g_{ab}T, \label{eq:2.10} \end{eqnarray} and it is this form of the field equations that is used here. The stress tensor is taken to vanish or to be that of a perfect fluid, \begin{equation} S_{ab}=(\mu+\nu)U_aU_b+\frac{1}{2}(\mu-\nu)g_{ab}, \label{eq:2.11} \end{equation}} %\indent where $\mu$ and $\nu$ are respectively the density and the pressure of the fluid. $U_a$ is a unit timelike vector field, \begin{equation} U_aU^a=-1, \label{eq:2.12} \end{equation}} %\indent with acceleration, expansion and projection, \begin{eqnarray} A^a&=&U^a_{~;b}u^b=\dot{U}^a,\nonumber\\ \Theta&=&u^a_{~;a},\\ i^{ab}&=&g^{ab}+U^aU^b,\nonumber \label{eq:2.13} \end{eqnarray} respectively. The first and second conservation equations are \begin{eqnarray} \mu_aU^a+(\mu+\nu)\Theta&=&0,\nonumber\\ (\mu+\nu)A^a+i^{ab}\nu_b&=&0, \label{eq:2.14} \end{eqnarray} respectively. \section{Poynting Vectors} \label{sec:PV} In a nonvacuum space-time the speed of a gravitational wave can be measured relative to the matter present. For example, in a cosmological model it can be compared to the speed of the comoving fluid velocity vector; on a galactic scale the speed of a gravitational wave emitted from the center of the galaxy could be estimated be comparing it to the speed of co-occurring electromagnetic radiation, or possibly by direct methods. The metrics that will be here have the property \ref{eq:6.1}, where the first conservation implies that the density of a fluid is at rest in this coordinate system, so that there is a possiblity of measuring the speed of a gravitational wave with respect to this fluid. So far the field equations coupled to a perfect fluid have proved intractable, and only solutions with zero stress have been found; thus for a timelike wave it is possible to move to a coordinate system where the wave appears to be at rest, and as there is no matter present this criterion cannot be used to measure wave speed. Another criterion is needed and the one suggested here is to construct Poynting vectors representing the speed of energy transfer: whether the Poynting vector is timelike, null or spacelike will be taken to support the belief that the gravitational wave is timelike, null or spacelike. If the principle of equivalence holds, then the energy of the gravitaional field appears to be nonexistent to an observer in free fall, and consequently a Poynting vector for gravitaional energy gives at best only an indication of the real energetics involved. To sum up: whether a gravitational enery Poynting vector is timelike, null or spacelike can be regarded as measuring whether the gravitational wave is timelike, null or spacelike; alternatively it can be considered as a measure of the appropriatenees, or otherwise, of the Poynting vector. There are tensors constructed out of products and derivatives of the Riemann tensor, which can represent the square of gravitational energy \cite{bi:bel} There are a large number of possible combinations \cite{bi:collinson}. An often-used combination is produced by first Matt\'e-decomposing the Weyl tensor \begin{eqnarray} C_{abcd}=R_{abcd}&+&\frac{1}{6}R(g_{ac}g_{db}-g_{ad}g_{cb})\nonumber\\ &+&\frac{1}{2}(g_{ad}R_{cb}-g_{ac}R_{db}+g_{bc}R_{da}-g_{db}R_{ca}), \label{eq:3.1} \end{eqnarray} into its electric and magnetic parts, \begin{equation} E_{ab}\equiv C_{ambn}U^mU^n,~~~ B_{ab}\equiv *C_{ambn}U^mU^n, \label{eq:3.2} \end{equation}} %\indent where $U_a$ is a timelike vector field and $*C_{abcd}=\frac{1}{2}\sqrt{-g}\epsilon_{abcd}C^{mn}_{..cd}$, and then defining \begin{equation} T_{abcd}\equiv C_{ambn}C^{~m~n}_{c.d.}+*C_{ambn}*C^{~m~n}_{c.d.}, \label{eq:3.3} \end{equation}} %\indent as an energy squared tensor, to give the Poynting vector \begin{eqnarray} P^a&\equiv&T^a_{.bcd}U^bU^cU^d\nonumber\\ &=&C^a_{.mbn}E^{mn}_{..}U^b+*C^a_{.mbn}B^{mn}_{..}U^b. \label{eq:3.4} \end{eqnarray} Energy tensors of the correct dimension can be constructed using the Lanczos tensor \cite{bi:mdr88}. Again there are a large number of possibilities, with the added complication of the choice of gauges possible for the Lanczos tensor. Here it is assumed that both the Lanczos algebraic \begin{equation} 3\chi_a\equiv H^{~b}_{a.b}=0, \label{eq:3.5} \end{equation}} %\indent gauge and differential \begin{equation} D_{ab}\equiv H^{~~c}_{ab.;c}, \label{eq:3.6} \end{equation}} %\indent gauge ave been applied. Subject to these gauges the Weyl tensor is recovered as a linear diferential equation in the Lanczos tensor \begin{eqnarray} C_{abcd}&=&H_{abc;d}-H_{abd;c}+H_{cda;b}-H_{cdb;a}\nonumber\\ &-&g_{ac}H^{~e}_{b.d;e}+g_{ad}H^{.e}_{b.dc;e} -g_{bd}H^{~e}_{a.c;e}+g_{bc}H^{~e}_{a.d;e}. \label{eq:3.7} \end{eqnarray} The Lanczos tensor has the symmetries \begin{eqnarray} &&H_{abc}=-H_{bac},\nonumber\\ &&H_{abc}=H_{cba}-H_{cab}. \label{eq:3.8} \end{eqnarray} Two of the more successfu; energy tensors are \begin{eqnarray} L_{ab}&=&H^{~cd}_{a..}H_{bcd}+*H^{~cd}_{a..}*H_{bcd}\nonumber\\ &=&2H^{~cd}_{a..}H_{bcd}-\frac{1}{2}g_{ab}H_{cde}H^{cde}_{...},\\ \label{eq:3.9} M_{ab}&=&2H^{~c}_{a.b;c}. \label{eq:3.10} \end{eqnarray} Their Poynting vectors can be defined as \begin{eqnarray} O^a_.&\equiv&L^{ab}_{..}U_b,\nonumber\\ Q^a_.&\equiv&M^{ab}_{..}U_b, \label{eq:3.11} \end{eqnarray} where $U_a$ is a timelike vector. \section{The Metric} \label{sec:metric} The line element is taken to be of the form \begin{equation} ds^2=-dt^2+dx^2+dy^2+dz^2+h(u,x,y)du^2, \label{eq:4.1} \end{equation}} %\indent where \begin{equation} u\equiv lt+z. \label{eq:4.2} \end{equation}} %\indent For $l=1$, $u$ is null and this is Peres' metric; for $l^2<1$ or $l^2>1$, $u$ is spacelike or timelike respectively. For $l^2<1$ define \begin{equation} \sqrt{1-l^2}\cdot z'\equiv lt+z\equiv u,~~~ \sqrt{1-l^2}\cdot t'\equiv t+lz. \label{eq:4.3} \end{equation}} %\indent Then the metric becomes \begin{equation} ds^2=-dt^2+dx^2+dy^2+[1+(1-l^2)h(z',x,y)]dz'^2, \label{eq:4.4} \end{equation}} %\indent which is manifestly static. For $l^2>1$ define \begin{equation} \sqrt{l^2-1}\cdot t''\equiv lt+z\equiv u,~~~ \sqrt{l^2-1}\cdot z''\equiv t+lz. \label{eq:4.5} \end{equation}} %\indent Then the metric becomes \begin{equation} ds^2=-[1+(1-l^2)h(\sqrt{l^2-1}\cdot t'',x,y)]dt''^2+dx^2+dy^2+dz''^2. \label{eq:4.6} \end{equation}} %\indent This metric has the lapse $N$ as the only varying component of the metric and may be written in the form \begin{equation} ds^2=-N^2dt^2+dx^2+dy^2+dz^2. \label{eq:4.7} \end{equation}} %\indent When the lapse is of the variables seperable form \begin{equation} N=H(t)\alpha(x,y), \label{eq:4.8} \end{equation}} %\indent the time dependence can be absorbed into the metric, showing that the metric is static. \section{The Field Equations} \label{sec:fe} The line element \ref{eq:4.7} gives the Christoffel symbols \begin{equation} \verb+{+^t_{tt}\verb+}+=\frac{\dot{N}}{N},~~~ \verb+{+^t_{ti}\verb+}+=\frac{N_i}{N},~~~ \verb+{+^i_{tt}\verb+}+=NN^i_., \label{eq:5.1} \end{equation}} %\indent where $i,j=x,y,z$ and $\dot{N}=\partial-tN$. The Riemann tensor, Ricci tensor and Ricci scalar have nonvanishing components: \begin{eqnarray} R^t_{.itj}&=&-\frac{N_{ij}}{N},~~~R^i_{.tjt}=NN^i_{.j},\label{eq:5.2}\\ R_{tt}&=&NN^{~i}_{i.},~~~R_{ij}=-\frac{N_{ij}}{N},\label{eq:5.3}\\ R&=&-2\frac{N^{~i}_{i.}}{N}.\label{eq:5.4} \end{eqnarray} The first covariant derivatives of the Riemann tensor are \begin{eqnarray} R_{itjt;t}&=&N^2\left(\frac{N_{ij}}{N}\right)^\circ,~~~ R_{itjt;k}=N^2\left(\frac{N_{ij}}{N}\right)_k,\label{eq:5.5}\\ R_{jtik;t}&=&N_iN_{jk}-N_kN_{ji}.\nonumber \end{eqnarray} The first covariant derivatives of the Ricci tensor can be found by contraction; note in particular \begin{equation} R_{it;t}=N^j_.N_{ij}-N_iN^{~j}_{j.}. \label{eq:5.6} \end{equation}} %\indent The second covariant derivatives of the Riemann tensor are \begin{eqnarray} R_{itjt;tt}&=&N^3 \left\{\left[\frac{1}{N}\left(\frac{N_{ij}}{N}\right)^\circ\right]^\circ -\left[N\left(\frac{N_{ij}}{N}\right)_k\right]^k\right\}\nonumber\\ &&+N^4\left(\frac{N_{ij}}{N}\right)^{~k}_{k.} -NN^k(R_{jtik;t}+R_{itjk;t}),\nonumber\\ R_{jtik;tt}&=&N^2\left[\frac{1}{N^2}(n_iN_{jk}-N_kN_{ji}\right]^\circ\nonumber\\ &&-N\left[N_i\left(\frac{N_{jk}}{N}\right)^\circ +N_k\left(\frac{N_{jk}}{N}\right)^\circ\right],\\ R_{ijkl;tt}&=&\frac{2}{N}(N-iN-kN_{jl}-N_iN_lN_{jk} -N_jN_kN_{ij}+N_lN_jN_{ik}),\nonumber\\ R_{itjt;kl}&=&N^2\left(\frac{N_{ij}}{N}\right)_{kl}.\nonumber \label{eq:5.7} \end{eqnarray} The tensors \ref{eq:2.9} and \ref{eq:2.6} are \begin{eqnarray} Z_{tt}=&2&\left(p-\frac{2b}{3}\right)N \left[Ng^{ab}\left(\frac{N^{~i}_{i.}}{N}\right)_a\right]_b -\frac{4bN^i_.}{N}R_{it;t}\nonumber\\ &+&4\left(p+\frac{b}{3}\right)NN^i_. \left(\frac{N^{~j}_{j.}}{N}\right)_i,\nonumber\\ Z_{it}=&-&2b\left[-N\left(\frac{R_{it;t}}{N^3}\right)^\circ+\frac{1}{N^2} (N_i\dot{N}^{~j}_{j.}-N^j_.\dot{N}_{ij})\right]\nonumber\\ &-&4\left(p+\frac{b}{3}\right)N \left[\frac{1}{N}\left(\frac{N^{~j}_{j.}}{N}\right)^\circ\right]_i\\ Z_{ij}=&-&2b\left\{\frac{1}{N}\left[Ng^{ab} \left(\frac{N_{ij}}{N}\right)_a\right]_b -\frac{1}{N^3}(N_iR_{jt;t}+N_jR_{it;t})\right\}\nonumber\\ &+&\left(p+\frac{b}{3}\right)\left\{ -4\left(\frac{N^{~k}_{k.}}{N}\right)_{ij} -\eta_{ij}\frac{2}{N}\left[Ng^{ab} \left(\frac{N^{~k}_{k.}}{N}\right)_a\right]_b\right\}, \label{eq:5.8} \end{eqnarray} and \begin{eqnarray} X_{tt}&=&3bN_{ij}N^{ij}_{..} +\left(2p-\frac{7b}{3}\right)N^{~i}_{i.}N^{~j}_{j.},\\ X_{it}&=&0,\nonumber\\ X_{ij}&=&-4p\left(p+\frac{b}{3}\right)\frac{N_{ij}N^{~k}_{k.}}{N^2} +\frac{\eta_{ij}}{N^2}\left(bN_{kl}N^{kl}_{..} +\left(2p-\frac{b}{3}\right)n^{~l}{l.}N^{~k}_{k.}\right),\nonumber \label{eq:5.9} \end{eqnarray} respectively. \section{Solutions to the Field Equations} \label{sec:sol} From vacuum general relativity, comparing the Ricci tensor \ref{eq:4.3} and the Riemann tensor \ref{eq:4.2} we see that the only solution is flat space. For a perfect fluid the first and second conservation equations \ref{eq:2.14} become \begin{eqnarray} \mu^0&=&0,\nonumber\\ (\mu+\nu)\frac{N_i}{N}&+&\nu_i=0, \label{eq:6.1} \end{eqnarray} respectively. The stress is given by \begin{eqnarray} S^{~t}{t.}&=&-\frac{1}{2}(3\mu+\nu),\nonumber\\ S_{ij}&=&\frac{1}{2}\eta_{ij}(\mu-\nu). \label{eq:6.2} \end{eqnarray} For the perfect fluid in general relativity, the $S_{xy}, S_{xz}, S_{zy}$ equations show that $N$ is a function of one spatial coordinate, say $x$, and then comparing $S_{xx}$ and $S_{yy}$ shows that this vanishes; again the only solution is flat space. In general the firld equations constructed in Sec \ref{sec:fe} are too complex to solve, especially because of the occurrence of $R_{it;t}$. Taking \begin{equation} x=r~sin \theta,~~~ y=r~cos \theta, \label{eq:6.3} \end{equation}} %\indent gives \begin{eqnarray} &&R_{rt;t}=\frac{1}{r^2}\left(N_\theta N_{\theta,r} -\frac{N^2_\theta}{rN}-N_rN_{\theta,\theta} +2rN^2_r\right) +N_zN_{rz}-N_rN_{zz},\nonumber\\ &&R_{\theta t;t}=N_rN_{\theta,r}-\frac{N_rN_\theta}{rN}-N_\theta N_{,rr} +N_zN_{\theta z}-N_\theta N_{zz},\\ &&R_{zt;t}=N_rN_{zr}-N_zN_{rr} +\frac{1}{r^2}[N_\theta N_{z\theta}-N_z(N_{\theta\th}-2rN_r)].\nonumber \label{eq:6.4} \end{eqnarray} Requiring $N=N(t,r)$ still leaves $R_{rt;t}$ nonvanishing. $R_{it;t}$ can be made to vanish by assuming that \begin{equation} N=N(t,k-ix^i). \label{eq:6.5} \end{equation}} %\indent Then if one lets $k^2=k_ik^i$ the field equations become \begin{eqnarray} S^{~t}_{.t}&=&-\frac{\Lambda}{\kappa^2}-\frac{k^2N''}{\kappa^2N} -\frac{2}{N}\left(p-\frac{2b}{3}\right) \left[Ng^{ab}\left(\frac{N''}{N}\right)_a\right]_b\nonumber\\ && +2k^2\left(p+\frac{b}{3}\right) \left\{\frac{2}{N}\left[\frac{1}{N}\left(\frac{N''}{N}\right)^\circ\right]^\circ -\frac{k^2}{N^2}\left[N''^2 +2NN'\left(\frac{N''}{N}\right)'\right]\right\},\nonumber\\ S^{~i}_{t.}&=&-4k^2K^i_.\left(p+\frac{b}{3}\right) N\left[\frac{1}{N}\left(\frac{N''}{N}\right)^\circ\right]',\\ S^{~j}_{i.}&=&k_ik^j_.\left\{-\frac{N''}{\kappa^2N} +\frac{2b}{N}\left[Ng^{ab}\left(\frac{N''}{N}\right)_a\right]_b -4k^2\left(p+\frac{b}{3}\right)\left[\left(\frac{N''}{N}\right)'' +\frac{N''^2}{N^2}\right]\right\}\nonumber\\ &&+2\left(p+\frac{b}{3}\right)\delta^{~j}_{i.}k^2\left\{-\frac{1}{N} \left[Ng^{ab}\left(\frac{N''}{N}\right)_a\right]_b+\frac{k^2N''^2}{N^2}\right\} -\frac{\Lambda}{\kappa^2}\delta^{~j}_{i.}.\nonumber \label{eq:6.6} \end{eqnarray} First consider the case \begin{equation} p+\frac{b}{3}=\frac{1}{6\kappa^2}\left(\frac{1}{m^2_{WP}}-\frac{1}{m^2_{RB}}\right)=0, \label{eq:6.7} \end{equation}} %\indent and $\Lambda=0$; then the field equations become \begin{eqnarray} S_{ab}&=&-2b\left(\Box-\frac{1}{2b\kappa^2}\right)R_{ab}\nonumber\\ &=&\frac{1}{\kappa^2m^2_{RB}}(\Box+m^2_{RB})R_{ab}. \label{eq:6.8} \end{eqnarray} For $S_{ab}=0$, by the use of \ref{eq:5.3} and \ref{eq:5.7} this equation is \begin{equation} 0=\left[\frac{1}{N}\left(\frac{N_{ij}}{N}\right)^\circ\right]^\circ -\left[N\left(\frac{N_{ij}}{N}\right)_k\right]^k -m^2_{RB}N_{ij}, \label{eq:6.9} \end{equation}} %\indent where $\dot{N}=\partial_tN$. This has the static solution \begin{equation} N=A(t)\exp\left(a_0+a_1z-\frac{1}{4}m^2_{RB}z^2\right), \label{eq:6.10} \end{equation}} %\indent where $a_0$ and $a_1$ are constants. This solution can be rotated around the spatial axis, and also transferred into the form \ref{eq:4.1}. A nonstatic solution can be found by assuming that the lapse $N$ is a function of $w=k_ax^a$, where $k_a$ is a constant. Symbolically this is \begin{equation} N=N(-k_0t+k_ix^i)=N(k_ax^a)=N(w). \label{eq:6.11} \end{equation}} %\indent The Riemann tensor takes the simple form \begin{equation} R^t_{.itj}=-k_ik_j\frac{N''}{N}. \label{eq:6.12} \end{equation}} %\indent If one defines \begin{equation} t\equiv u=\frac{lt'+z'}{\sqrt{l^2-1}},~~~ z\equiv\frac{t'+lz'}{\sqrt{l^2-1}}, \label{eq:6.13} \end{equation}} %\indent the metric can be put in the form \ref{eq:3.1} with \begin{equation} h=1-N^2\left[u\left(-k_0+\frac{k_z}{l}\right)+ k_xx+k_yy+k_z\frac{\sqrt{l^2-1}}{l}z'\right]. \label{eq:6.14} \end{equation}} %\indent If one assumes that $p=-b/3$ as in \ref{eq:6.7}, the field equations are \begin{eqnarray} S^{~t}_{t.}=S^{~i}_{i.},~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~\nonumber\\ \kappa^2m_2^2S_{ij}=k_ik_j\left\{-m^2_2\frac{N''}{N} -\frac{k_0^2}{N}\left[\frac{1}{N}\left(\frac{N''}{N}\right)'\right]' +\frac{k^2}{N}\left[N\left(\frac{N''}{N}\right)'\right]'\right\}. \label{eq:6.15} \end{eqnarray} When $S_{ab}=0$ these equations reduce to the single equation \begin{equation} -\frac{k^2_0}{N}\left[\frac{1}{N}\left(\frac{N''}{N}\right)'\right]' +\frac{k^2}{N}\left[N\left(\frac{N''}{N}\right)'\right]'-m^2_2\frac{N''}{N}=0. \label{eq:6.16} \end{equation}} %\indent Integrating once, with the constant of integration $a_1$, and multiplying by $N$ gives \begin{equation} (k^2N^2-k_0^2)\left(\frac{N''}{N}\right)'-m_2^2NN'-a_1N=0. \label{eq:6.17} \end{equation}} %\indent Setting the constant of integration $a_1=0$ and integrating again with the constant of integration $a_2$ gives \begin{equation} \frac{N''}{N}=a_2+\frac{1}{2}m_2^2k^{-2}\ln(k^2N^2-k_0^2). \label{eq:6.18} \end{equation}} %\indent This differential equation remains intractable when $m_2\ne0$, but when $m_2=0$ it has the simple solution \begin{equation} N=C_+\exp(k_ax^a)+C_-\exp(-k_ax^a), \label{eq:6.19} \end{equation}} %\indent where a factor of $\sqrt{a_2}$ has been absorbed into $k_a$, and thus $\sqrt{a_2}k_a\rightarrow k_a$; because $a_2$ is a constant of integration $k_a$ may now be complex. For \ref{eq:6.19} the Riemann tensor takes the simple form \begin{equation} R^t_{.itj}=-k_ik_j. \label{eq:6.20} \end{equation}} %\indent For $p+\frac{b}{3}\ne0$, inspection of the $S^{~i}_{t.}$ term \ref{eq:6.6} and the fluid conservation equations \ref{eq:6.1} suggests that it should be possible to integrate the field equations, but so far this has turned out not to be the case. It is known \cite{bi:mdr89} that a sperically symmetric metric is always a solution of the general relativity field equations if a sufficiently large number of other fields and fluids are present. Adding more terms to the stress in \ref{eq:6.6} has not yet given invertible equations. To find $p+\frac{b}{3}\ne0$ solutionsthe expresion \ref{eq:6.20}, i.e. $R^t_{.itj}=-k_ik_j=-N_{ij}/N$, is substituted back into the field equations \ref{eq:6.6}. This assumption amounts to disregarding higher derivatives of the Rieman tensor and investigating only the effect of cross terms; therefore it complements previous linear analysis. Substituting \ref{eq:6.6} gives the field equations \begin{eqnarray} S^{~t}_{t.}&=&-\frac{k^2}{\kappa^2} -\frac{\Lambda}{\kappa^2} -2k^4\left(p+\frac{b}{3}\right),\nonumber\\ S^{~i}_{t.}&=&0,\\ S^{~i}_{i.}&=&-\frac{k^2}{\kappa^2} -\frac{3\Lambda}{\kappa^2} +2k^4\left(p+\frac{b}{3}\right),\nonumber\\ S^{~j}_{i.}(i\ne j)&=&k_ik^j\left[-\frac{1}{\kappa^2} -4k^2\left(p+\frac{b}{3}\right)\right].\nonumber \label{eq:6.21} \end{eqnarray} Taking $S_{ab}=0$ gives the solution \begin{eqnarray} k^2&=&-2\Lambda=\frac{-1}{4\kappa^2\left(p+\frac{b}{3}\right)}\nonumber\\ &=&\frac{-3}{2(m^{-2}_{WP}-m^{-2}_{RB})}. \label{eq:6.22} \end{eqnarray} \section{Poynting Vectors for the Metric} \label{sec:Poy} From Eqs. \ref{eq:3.1} and \ref{eq:4.2} the Weyl tensor is found to be \begin{eqnarray} C^t_{.itj}&=&-\frac{N_{ij}}{2N}+\frac{N^{~k}_{k.}}{6N}\eta_{ij},\nonumber\\ C^i_{jkl}&=&\frac{1}{2N}(-\delta^i_lN_{jk}+\delta^i_kN_{jl} -\eta_{jk}N^i_l+\eta_{jl}N^i_k)\\ &&-\frac{N^{~m}_{m.}}{3N}(\delta^i_k\eta_{jl}-\delta^i_l\eta_{jk}).\nonumber \label{eq:7.1} \end{eqnarray} The electric and magnetic parts become \begin{eqnarray} E_{ij}&=&\frac{N_{ij}}{2N}-\frac{N^{~k}_{k.}}{6N}\eta_{ij},\nonumber\\ B_{ab}&=&0. \label{eq:7.2} \end{eqnarray} The Poynting vector \ref{eq:3.4} is \begin{equation} P^a_.=\frac{\delta^a_t}{12N^3}(3N_{ij}^{~~2}-N_{ii}^{~~2}), \label{eq:7.3} \end{equation}} %\indent and thus \begin{equation} P_aP^a=\frac{-1}{12^2N^4}(3N_{ij}^{~~2}-n_{ii}^{~~2})^2, \label{eq:7.4} \end{equation}} %\indent is always timelike. This Poynting vector appears to give no indication of energy transfer as $P^t_.$ is the only nonvanishing component. The Lanczos tensor can be calulated by direct methods similar to those used in Ref.\cite{bi:mdr95}. In the Lanczos algebraic and differential gauges it is found to be \begin{equation} H^t_{.it}=-\frac{N_i}{3N},~~~ H_{ijk}=\frac{1}{6N}(N_j\eta_{ik}-N_i\eta_{jk}). \label{eq:7.5} \end{equation}} %\indent The energy tensor \ref{eq:3.9} is \begin{equation} L_{tt}=-\frac{N_i^2}{18},~~~ L_{it}=0,~~~ L_{ij}=\frac{1}{18N^2}(5N_iN_j-2\eta_{ij}N^2_k), \label{eq:7.6} \end{equation}} %\indent and the associated Poynting vector is \begin{equation} O^a_.=-\frac{N^2_i\delta^a_t}{18N^2}; \label{eq:7.7} \end{equation}} %\indent thus \begin{equation} O_aO^a=-\left(\frac{N_i^2}{18N^2}\right)^2, \label{eq:7.8} \end{equation}} %\indent and again this is always timelike. The energy tensor \ref{eq:3.10} is \begin{eqnarray} M^t_{.t}&=&-\frac{2}{3}\left(\frac{N^i_.}{N}\right)_i,~~~ M^i_{.t}=-\frac{2}{3}\left(\frac{N^i_.}{N}\right)^\circ,\\ M_{ij}&=&\frac{2N_iN_j}{3N^2}-\frac{N_{ij}}{3N}+\eta_{ij}\frac{N^k_{.k}}{3N},\nonumber \label{eq:7.9} \end{eqnarray} and the associated Poynting vector is \begin{equation} Q^a_.=\frac{2}{3N}\left[ \left(\frac{N^i_.}{N}\right)_i,\left(\frac{N^i_.}{N}\right)^\circ\right], \label{eq:7.10} \end{equation}} %\indent giving \begin{equation} Q_aQ^a=-\frac{4}{9}g^{ab}\left(\frac{N^i_.}{N}\right)_a\left(\frac{N_i}{N}\right)_b. \label{eq:7.11} \end{equation}} %\indent For \ref{eq:6.20} with either $C_+=0$ or $C_-=0$, $Q^a=0$; however, for $C=C_+=C_-$, \begin{equation} Q_aQ^a=-\frac{4}{9}k^2\left(k^2-\frac{k_0^2}{N^2}\right)Sech^4(k_ax^a). \label{eq:7.12} \end{equation}} %\indent For the solution \ref{eq:6.22}, \ref{eq:7.12} reduces to \begin{equation} Q_aQ^a=\frac{-1}{(m^{-2}_{WP}-m^{-2}{_RB})} \left(\frac{1}{m^{-2}_{WP}-m^{-2}_{RB}}+\frac{2k_0^2}{3N^2}\right) \cdot Sech(k_ax^a). \label{eq:7.13} \end{equation}} %\indent When $m_{RB}=0$, $Q^a$ is always timelike; when $M_{RB}\ne 0$ a variety of behaviour is possible, depending on the values of the constants. \section{Acknowledgements} I would like to thank the Leverhulme Trust for financial support.
\section{Introduction} \noindent In the last years, the prediction and the observation of many microlensing events are gathering ever more interest in gravitational lensing. The typical light amplification curve for these events, found by Paczynski in 1986, has been observed by several astronomical collaborations in observation campaigns toward the bulge of our galaxy (Udalski et al., 1993a; Alard et al., 1997), the Large Magellanic Cloud and the Small Magellanic Cloud (Alcock et al., 1993; Aubourg et al., 1993), the spiral arms and Andromeda galaxy (Tomaney \& Crotts, 1996; Ansari et al., 1997; Melchior et al., 1999). Recently, observations toward globular clusters have even been suggested (Jetzer, Strassle \& Wandeler, 1998). Beyond proving the correctness of Paczynski's predictions, the observation of microlensing events provides a very cunning instrument for the investigation of the halo of our (and/or some other) galaxy. Together with events strictly following Paczynski's curve, some events showing deviations from the standard behaviour have been detected. Each of these deviations has found some interpretation (Finite source cut-off (Witt \& Mao, 1994; Alcock et al., 1998), blending (Sutherland, 1998), parallax effect (Gould, 1992; Alcock et al., 1995), binary lens (Schneider \& Weiss, 1986; Mao \& Paczynski, 1991; Udalski et al., 1993b). Indeed, the most intriguing of these deviations is the one induced by a binary (or multiple) lens. The study of light amplification curves produced by multiple lenses has not yet been performed analytically because of the difficulties in the inversion of the lens application. Anyway, these curves can be obtained numerically by using some inversion algorithm (inverse ray shooting)(Wambsganns, 1997). From the analytical point of view, only the caustics of a general binary lens have been studied in some detail (Witt \& Petters, 1993). The lack of analytical results utilizable in microlensing constitutes an irksome obstacle in the complete interpretation of multiple microlensing events. A particularly interesting case of multiple Schwarzschild lenses is formed when one of the masses is much biggest than others (Mao \& Paczynski, 1991; Gould \& Loeb, 1992). This is the situation of a typical planetary system where a central star is surrounded by its planets bearing masses thousand or million times smaller. The perturbations on Paczynski's curve induced by the presence of a (even Earth - like) planet are in principle detectable by collaboration teams exploiting world wide telescopic networks (Peale, 1997; Sackett, 1997). Then microlensing could become a new efficient method for the detection of small planets in extra - solar systems. This justifies the major interest in this field that is growing in the last months. Preliminary calculations on the probability of detection of planets have been made (Gould \& Loeb, 1992; Bolatto \& Falco, 1994) and great efforts are lavished on the problem of extraction of planetary parameters by approximate models (Gaudi \& Gould, 1996). It is easy to imagine how the availability of an analytical expression for light curves could help the researches in this field. The aim of this work is to describe planetary effects perturbatively, exploiting the very little ratios of the masses of the planets with respect to the star mass (Gould \& Loeb in 1992 first pioneered this kind of approach). In section 2 the lens equation and other usual objects are specified for the case of planetary systems. In section 3, by means of perturbative theory, I derive the complete structure of critical curves and caustics of a general (not only binary) planetary system; position of planetary caustics and cusps are also found. In section 4 the problem of the inversion of lens application is faced and resolved; consequently, analytical microlensing light curves for planetary events thus obtained are shown. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{F5VDNX0T.eps}} \caption{Tipical scheme of planetary lensing.} \label{Fig general lensing} \end{figure} \section{Planetary lensing} \noindent A planetary system is nothing but a discrete set of Schwarzschild lenses in a small portion of the space. Fig. \ref{Fig general lensing} shows the typical situation of planetary lensing. The lens plane is defined as a plane orthogonal to the line of sight situated on the barycentre of the lens mass distribution. According to the ordinary theory of lensing (Schneider, Ehlers \& Falco, 1992), if the scale of this distribution is much smaller than the distances separating the lens from the source and the observer, then one can deal with the density projected on the lens plane instead of considering the original volume density. This hypothesis is quite verified in real observations where the typical distances are at least of the order of kpc. This consideration can be important in planetary systems where very far planets can eventually have projections enough near to the central star to give perturbations comparable to those of planets placed in more favourable positions. This means that multiple events could be less out of common than one could think (Wambsganns, 1997; Gaudi, Naber \& Sackett, 1998). So, it is desirable to preserve the whole planetary system as much as possible before abandoning it for the simplest case of the single planet around the big star. We shall see that the perturbative theory has the considerable advantage of being rather insensitive to the number of planets as regards the difficulty of the problem. Let's define the length $R_{E}^{\odot }=\sqrt{\frac{4GM_{\odot }}{c^{2}}% \frac{D_{LS}D_{OL}}{D_{OS}}}$. $\mathbf{x}=\left( x_{1};x_{2}\right) $ shall denote the coordinates in the lens plane normalized to $R_{E}^{\odot }$% , while $\mathbf{y}=\left( y_{1};y_{2}\right) $ shall be the coordinates in the source plane normalized to $R_{E}^{\odot }\frac{D_{OS}}{D_{OL}}$. $% m_{1}$ will be the mass of the central star and $m_{2}$, ..., $m_{n}$ will be the masses of the planets. All of these masses are meant to be measured in solar masses. The star will always be placed at the origin, while the projection on the lens plane of the i$_{th}$ planet will be denoted by $% \mathbf{x}_{i}=\left( x_{i1};x_{i2}\right) $. With these notations, the lens equation reads: \begin{equation} \mathbf{y}=\mathbf{x}-\frac{m_{1}\mathbf{x}}{\left| \mathbf{x}% \right| ^{2}}-\sum\limits_{i=2}^{n}\frac{m_{i}\left( \mathbf{x}-% \mathbf{x}_{i}\right) }{\left| \mathbf{x}-\mathbf{x}_{i}\right| ^{2}% } \label{General lens equation} \end{equation} In (\ref{General lens equation}), the deviation of light rays due to the star has been explicitly separated by those of the planets. Given a source position $\mathbf{y}$, the corresponding $\mathbf{x}$ solving the lens equation are called images. Many interesting properties of this vectorial application can be studied through its Jacobian matrix. In particular, the determinant of this matrix contains nearly all the information about the properties of the images. Let's write the Jacobian determinant for the case of planetary lensing using (\ref{General lens equation}): \begin{multline} \det J=1-\left[ \frac{m_{1}\left( x_{1}^{2}-x_{2}^{2}\right) }{\left( x_{1}^{2}+x_{2}^{2}\right) ^{2}}+\sum\limits_{i=2}^{n}\frac{m_{i}\left( \Delta _{i1}^{2}-\Delta _{i2}^{2}\right) }{\left( \Delta _{i1}^{2}+\Delta _{i2}^{2}\right) ^{2}}\right] ^{2}+ \label{General Jacobian} \\ -4\left[ \frac{m_{1}x_{1}x_{2}}{\left( x_{1}^{2}+x_{2}^{2}\right) ^{2}}% +\sum\limits_{i=2}^{n}\frac{m_{i}\Delta _{i1}\Delta _{i2}}{\left( \Delta _{i1}^{2}+\Delta _{i2}^{2}\right) ^{2}}\right] ^{2} \end{multline} where $\mathbf{\Delta }_{i}=\left( \Delta _{i1};\Delta _{i2}\right) =% \mathbf{x}-\mathbf{x}_{i}.$ Given an image I at position $\mathbf{x}% _{I}$, the sign of $\det J\left( \mathbf{x}_{I}\right) $ is called the parity of I. It can be proved that the amplification of the image I is given by \begin{equation} \mu _{I}=\frac{1}{\left| \det J\left( \mathbf{x}_{I}\right) \right| } \label{General amplification} \end{equation} We see from this equation that when $\det J$ is null, the amplification diverges. This is rigorously true for point sources (in ray optics), while for finite (real) sources the integration over the source's surface makes the amplification finite (Witt \& Mao, 1994). The points where $\det J$ vanishes are called critical points, the corresponding points in the source plane through (\ref {General lens equation}) constitute the caustics. So, a point source crossing a caustic will produce images with infinite amplification. Real sources crossing caustics by some of their parts will be highly (but not infinitely) amplified. The structure of critical curves and caustics provides a substantial description of the general behaviour of the lens (number and roughly location of the images can be established). In microlensing, it is possible to give a qualitative description of light curves just checking whether the track of the source threads some caustic or not. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{F5VDNX0U.eps}} \caption{There are three possible situations in planetary lensing as regards critical curves, shown in these plots. The smooth round curves are critical curves, while the small cusped curves are the corresponding caustics.} \label{Fig general critical curves} \end{figure} In fig. \ref{Fig general critical curves} the (numerically obtained) critical curves and the caustics of a star with a single planet are shown. There are three possible cases in such a situation. I recall that for a single point source the critical curve is a ring with radius given by its Einstein radius $\sqrt{m}$, while the caustic is a point in the origin of the source plane. When the planet is far beyond the star's Einstein ring, there is only a small perturbation of the two rings which lends finite extension to the originally point - like caustics. This is much more evident in the planetary caustic which is also displaced towards the star. When the planet is in the proximity of the star's Einstein ring, the two critical curves merge and so do the caustics. In the last situation where the planet is internal to the star's Einstein ring, the star's critical curve returns to be very near to a ring while the planet's critical curve turns into two ovals to which a couple of triangular caustics correspond behind the star. \section{Caustics and perturbative analysis} \noindent In the solar system, the mass ratios between planets and the sun are always less than one thousandth. Jupiter is $9.4\times 10^{-4}M_{\odot }$% . Other planets are even less: Earth is $3\times 10^{-6}M_{\odot }$. With these numbers, it is natural to expect that the presence of planets should cause only little perturbations to the single lens case. Upon this consideration the perturbative hypothesis is based. In this and the following section the ratios between planetary and stellar masses will play the role of perturbative expansions parameters. We shall see that in most cases a first order expansion is sufficient to get very reliable results. Let's turn to the caustics of planetary systems. First I shall examine the modifications induced in the star's Einstein ring and consequently the central caustic. Then I shall deal with planetary caustics. \subsection{Central caustic} \begin{figure*} \resizebox{12cm}{!}{\includegraphics{F9I29S00.eps}} \hfill \parbox[b]{55mm}{ \caption{Central caustic for Jovian planets. The caustics on the left are those found perturbatively, while those on the right are obtained numerically. Little differences occur when one of the planets is close to the star's Einstein ring.} \label{Fig central caustic (Jupiter)}} \end{figure*} Of course, the starting point for the study of critical curves is the equation $\det J=0$, which can easily be written in polar coordinates: \begin{multline} \left[ \frac{m_{1}\left( \cos ^{2}\vartheta -\sin ^{2}\vartheta \right) }{% r^{2}}+\sum\limits_{i=2}^{n}\frac{m_{i}\left( \Delta _{i1}^{2}-\Delta _{i2}^{2}\right) }{\left( \Delta _{i1}^{2}+\Delta _{i2}^{2}\right) ^{2}}% \right] ^{2}+ \\ +4\left[ \frac{m_{1}\sin \vartheta \cos \vartheta }{r^{2}}% +\sum\limits_{i=2}^{n}\frac{m_{i}\Delta _{i1}\Delta _{i2}}{\left( \Delta _{i1}^{2}+\Delta _{i2}^{2}\right) ^{2}}\right] ^{2}=1 \end{multline} Here $\mathbf{\Delta }_{i}=\left( r\cos \vartheta -x_{i1};r\sin \vartheta -x_{i2}\right) $. Expanding this equation to the first order in $m_{i}$, we get: \begin{multline} \frac{m_{1}^{2}}{r^{4}}+\frac{2m_{1}\left( \cos ^{2}\vartheta -\sin ^{2}\vartheta \right) }{r^{2}}\sum\limits_{i=2}^{n}\frac{m_{i}\left( \Delta _{i1}^{2}-\Delta _{i2}^{2}\right) }{\left( \Delta _{i1}^{2}+\Delta _{i2}^{2}\right) ^{2}}+ \label{Central detJ=0} \\ +\frac{8m_{1}\sin \vartheta \cos \vartheta }{r^{2}}\sum\limits_{i=2}^{n}% \frac{m_{i}\Delta _{i1}\Delta _{i2}}{\left( \Delta _{i1}^{2}+\Delta _{i2}^{2}\right) ^{2}}=1 \end{multline} The zeroth order solution is simply $r=\sqrt{m_{1}}$, i.e. the Einstein ring. Let's write the first order solution as: \begin{equation} r=\sqrt{m_{1}}\left( 1+\varepsilon \left( \vartheta \right) \right) \end{equation} with $\varepsilon \ll 1$. Substituting in (\ref{Central detJ=0}) and expanding to the first order in $\varepsilon $. The zeroth order solution cancels and $\varepsilon $ is found solving the remaining first degree equation: \begin{multline} \varepsilon \left( \vartheta \right) =\frac{1}{2}\left( \cos ^{2}\vartheta -\sin ^{2}\vartheta \right) \sum\limits_{i=2}^{n}\frac{m_{i}\left[ \left( \Delta _{i1}^{0}\right) ^{2}-\left( \Delta _{i2}^{0}\right) ^{2}\right] }{% \left[ \left( \Delta _{i1}^{0}\right) ^{2}+\left( \Delta _{i2}^{0}\right) ^{2}\right] ^{2}}+ \label{Central critic curve} \\ +2\sin \vartheta \cos \vartheta \sum\limits_{i=2}^{n}\frac{m_{i}\Delta _{i1}^{0}\Delta _{i2}^{0}}{\left[ \left( \Delta _{i1}^{0}\right) ^{2}+\left( \Delta _{i2}^{0}\right) ^{2}\right] ^{2}} \end{multline} where $\mathbf{\Delta }_{i}^{0}=\left( \sqrt{m_{1}}\cos \vartheta -x_{i1};% \sqrt{m_{1}}\sin \vartheta -x_{i2}\right) $ is $\mathbf{\Delta }_{i}$ to the zeroth order. By very few steps we have found the perturbation of the Einstein ring in a very simple way. The parametric equation of the central caustic is soon found by applying the lens equation (\ref{General lens equation}) and expanding again to the first order in $m_{i}$: \begin{mathletters} \label{Central caustic} \begin{align} y_{1}\left( \vartheta \right) & =2\sqrt{m_{1}}\varepsilon \left( \vartheta \right) \cos \vartheta -\sum\limits_{i=2}^{n}\frac{m_{i}\Delta _{i1}^{0}}{% \left[ \left( \Delta _{i1}^{0}\right) ^{2}+\left( \Delta _{i2}^{0}\right) ^{2}\right] } \\ y_{2}\left( \vartheta \right) & =2\sqrt{m_{1}}\varepsilon \left( \vartheta \right) \sin \vartheta -\sum\limits_{i=2}^{n}\frac{m_{i}\Delta _{i2}^{0}}{% \left[ \left( \Delta _{i1}^{0}\right) ^{2}+\left( \Delta _{i2}^{0}\right) ^{2}\right] } \end{align} \end{mathletters} Of course, perturbative results are characterized by precise limits of validity. In our case we see that when the denominators in (\ref{Central critic curve}) vanish the perturbation diverges. This is not allowed by our assumption that $\varepsilon $ must be very small with respect to the unity. Those denominators represent the distance between the planet and the general point of the unperturbed star's Einstein ring. So we expect the perturbative theory to fail in those portions of the critical curve which are very near to one of the planets at least. We can understand this ``failure'' if we look back at fig. \ref{Fig general critical curves}b: when the planet is close to the star's Einstein ring, there is only one critical curve which is somewhat different from the ring in the proximity of the planet. For some values of $\vartheta $, the radial coordinate describing the critical curve assumes also more than one value; this situation cannot be described by a first order approximation, where, as we saw, the perturbation solves a first degree equation. The most interesting aspect of eqs. (\ref{Central critic curve}) and (\ref {Central caustic}) is that they are comprehensive of the action of the whole planetary system: they are valid for an arbitrary number of planets, not only the classically investigated case of the single planet. So these formulas enjoy a very high generality and can be used in more realistic contexts. We also note that the contributions coming from different planets superpose without interfering. This is an obvious consequence of the first order approximation; if I had included second order terms, I would have found ``interaction'' between planets. These ``interaction'' terms are thus not relevant in a first approximation. Now let's compare the perturbative caustics with those found by classical numerical algorithms to test the validity of the perturbative approach. In fig. \ref{Fig central caustic (Jupiter)} I show the results for the case of two Jovian planets placed in several positions. When the planets are far enough from the Einstein ring (fig. \ref{Fig central caustic (Jupiter)}a, \ref{Fig central caustic (Jupiter)}d), the caustic found according to (\ref {Central caustic}) is entirely identical to the numerical one. Letting one of the planets approach the Einstein ring, a small deviation begins appearing in the region coming from the portion of the Einstein ring that is closest to the planet. This deviation manifests itself in the size of the largest cusp. For Jovian planets these discrepancies unveil at distances from the star's Einstein ring of the order of a tenth of the Einstein radius. These first encouraging results become much better in the case of Earth - like planets. We expect the range of validity of perturbative results to be increased for this kind of planets, because of their smaller mass. Fig. \ref{Fig central caustic (Earth)} shows that for these little planets things go very well down to a hundredth of the Einstein radius. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{F9GZN001.eps}} \caption{Central caustics for Earth - like planets. The perturbative ones are on the left and the numerical on the right.} \label{Fig central caustic (Earth)} \end{figure} So the perturbative method is likely to provide reliable results at the first order already. Moreover, it is not to be forgotten that, in principle, the approximations can be improved pushing farther the perturbative expansion. \subsection{Planetary caustics} Planetary caustics are usually studied considering the planet as a point-lens with an external shear due to the star's gravitational field (Schneider, Ehlers \& Falco, 1992). This kind of lens was introduced by Chang \& Refsdal (1979; 1984) in a cosmological context. However, this lens is valid in planetary systems only to the lowest non-trivial order in $m_{i}$. Therefore a correct study should only retain the lowest order terms in the critical curves equation, so that Chang \& Refsdal's caustics are a suitable approximation at the lowest order only and not beyond. In order to complete the discussion of caustics in planetary systems and study their features properly, in this subsection I derive planetary caustics from perturbative hypothesis paying full attention to the order of each term. The situation for planetary caustics is rather different from that of the central caustic. There is no zeroth order solution to start from, since their very presence is perturbative. Nevertheless this is not a great problem: in fact we shall just search for the lowest order solution of the critical curves equation. To achieve this, I now rewrite $\det J=0$ in polar coordinates choosing the planet $m_{2}$ situated in $\mathbf{x}_{2}$ as the origin: \begin{multline} \left[ \frac{m_{1}\left( \Delta _{11}^{2}-\Delta _{12}^{2}\right) }{\left( \Delta _{11}^{2}+\Delta _{12}^{2}\right) ^{2}}+\frac{m_{2}\left( \cos ^{2}\vartheta -\sin ^{2}\vartheta \right) }{r^{2}}+ \right. \label{Planetary detJ=0} \\ \left. + \sum\limits_{i=3}^{n}% \frac{m_{i}\left( \Delta _{i1}^{2}-\Delta _{i2}^{2}\right) }{\left( \Delta _{i1}^{2}+\Delta _{i2}^{2}\right) ^{2}}\right] ^{2}+ 4\left[ \frac{m_{1}\Delta _{11}\Delta _{12}}{\left( \Delta _{11}^{2}+\Delta _{12}^{2}\right) ^{2}}+ \right. \\ \left. + \frac{m_{2}\sin \vartheta \cos \vartheta }{r^{2}}% +\sum\limits_{i=3}^{n}\frac{m_{i}\Delta _{i1}\Delta _{i2}}{\left( \Delta _{i1}^{2}+\Delta _{i2}^{2}\right) ^{2}}\right] ^{2}=1 \end{multline} When the planet is very far from the star, we know that its critical curve tends to an Einstein ring with radius $r=\sqrt{m_{2}}$. So we search for critical curves solving (\ref{Planetary detJ=0}) with $r\sim \sqrt{m_{2}}$. Let's save the lowest order terms only. In this operation, the contributions coming from the other planets are ejected out from the equation. It is convenient to place the star in the usual position $(-x_{p};0)$. What remains is: \begin{equation} \frac{m_{1}^{2}}{x_{p}^{4}}+\frac{m_{2}^{2}}{r^{4}}+\frac{2m_{1}m_{2}}{% x_{p}^{2}r^{2}}\left( \cos ^{2}\vartheta -\sin ^{2}\vartheta \right) =1 \end{equation} which is biquadratic in $r$. The solution is: \begin{equation} r=\sqrt{m_{2}\frac{\frac{m_{1}}{x_{p}^{2}}\cos 2\vartheta \pm \sqrt{1-\frac{% m_{1}^{2}}{x_{p}^{4}}\sin ^{2}2\vartheta }}{\left( 1-\frac{m_{1}^{2}}{% x_{p}^{4}}\right) }} \label{Planetary critical curve} \end{equation} which verifies our assumption $r\sim \sqrt{m_{2}}$. The parametric equations of caustics can be found in the usual way substituting (\ref{Planetary critical curve}) in the lens application and expanding to the first non trivial order ($\sqrt{m_{2}}$): \begin{mathletters} \label{Planetary caustics} \begin{eqnarray} y_{1} &=&x_{p}-\frac{m_{1}}{x_{p}}\left( 1-\frac{r}{x_{p}}\cos \vartheta \right) +\left( r-\frac{m_{2}}{r}\right) \cos \vartheta \\ y_{2} &=&-\frac{m_{1}}{x_{p}^{2}}r\sin \vartheta +\left( r-\frac{m_{2}}{r}% \right) \sin \vartheta \end{eqnarray} \end{mathletters} The contributions from the other planets are again of higher order. So the structure of planetary caustics is not affected by the presence of other planets at the lowest order in a perturbative expansion. These formulas can thus be used in a single planet situation as well as in a rich planetary system. Observe that $r$ goes to infinity as $x_{p}$ tends to $\sqrt{m_{1}}$, i.e. when the planet is next to the star's Einstein radius. So the perturbative results will not be valid in this situation. The reason is the same discussed for the central caustic. The merging of the two caustic is not describable in the lowest order perturbative expansion. Moreover, there's another limit to be taken in account. I have eliminated all the terms coming from the other planets because of their higher order. But these terms can become dominant when their denominators are small. This happens when one of these planets is close to the planet we are examining. This is not an exotic situation since we must always remember that what counts is actually the projection of the positions on the lens plane. So planets could be very far apart but have near projections yielding exotic critical curves. We see that the critical curves traced by (\ref{Planetary critical curve}) have two branches according to the double sign in their expression. For planets external to the star's Einstein ring ( $x_{p}>\sqrt{m_{1}}$), the branch coming from the positive sign is real while the other coming from the negative sign is imaginary for all values of $\vartheta $, being $\left| \frac{m_{1}}{x_{p}^{2}}\cos 2\vartheta \right| <\sqrt{1-\frac{m_{1}^{2}}{% x_{p}^{4}}\sin ^{2}2\vartheta }$. For internal planets ( $x_{p}<\sqrt{m_{1}}$% ), the denominator is negative and $\left| \frac{m_{1}}{x_{p}^{2}}\cos 2\vartheta \right| >\sqrt{1-\frac{m_{1}^{2}}{x_{p}^{4}}\sin ^{2}2\vartheta }$% . So the two branches are both real for: \begin{equation} \sin ^{2}2\vartheta <\frac{x_{p}^{4}}{m_{1}^{2}};\RIfM@\expandafter\text@\else\expandafter\mbox\fi{ \ }\cos 2\vartheta <0 \end{equation} that is in two small regions around $\vartheta =\frac{\pi }{2}$ and $% \vartheta =\frac{3\pi }{2}$. They are both imaginary elsewhere. All these results are coherent with the behaviour exposed in fig. \ref{Fig general critical curves}. We have one planetary caustic for external planets and two disconnected caustics for internal planets. \begin{figure*} \resizebox{12cm}{!}{\includegraphics{F5VAFT04.eps}} \hfill \parbox[b]{55mm}{ \caption{Planetary caustics for a Jovian planet. The caustics on the left column are perturbative while the ones on the right are numerical.} \label{Fig planetary caustics}} \end{figure*} Fig. \ref{Fig planetary caustics} shows the comparison with the numerical caustics. In fig. \ref{Fig planetary caustics}b the discrepancy with the numerical results appears as a loss of symmetry of the numerical caustic which is elongated towards the central star. This effect is not present in the perturbative Chang \& Refsdal's one. For internal planets near the star's Einstein ring, the basis of the triangular perturbative caustic is parallel to the star - planet axis(fig. \ref{Fig planetary caustics}c). So Chang \& Refsdal's lens works good until the field can be taken as uniform. When the spherical symmetry becomes important, the caustics begin to differ from perturbative ones. These effects can be taken into account by considering higher order terms in the expansion. These terms would provide the right corrections to the Chang \& Refsdal's approximation. Eqs. (\ref{Planetary critical curve}) and (\ref{Planetary caustics}) can be employed to find interesting characteristics of planetary caustics. For example, let's find the position of the couple of caustics for internal planets. We saw that the critical curves are centered upon $\vartheta =\frac{% \pi }{2}$ and $\vartheta =\frac{3\pi }{2}$. Consider the first of these (the other is similar at all). Inserting these values of $\vartheta $ in (\ref {Planetary critical curve}), the possible values of $r$ are: \begin{equation} r=\sqrt{m_{2}\frac{-\frac{m_{1}}{x_{p}^{2}}\pm 1}{\left( 1-\frac{m_{1}^{2}}{% x_{p}^{4}}\right) }} \end{equation} The point \begin{equation} r=x_{p}\sqrt{\frac{m_{1}m_{2}}{m_{1}^{2}-x_{p}^{4}}} \end{equation} obtainable by a quadratic mean from the two values, is internal to the critical curve and gives an approximation for its position. Immediately, using the lens equation and expanding to the lowest order terms, we find the position of the caustics: \begin{mathletters} \label{Planetary caustics position} \begin{eqnarray} y_{1} &=&x_{p}-\frac{m_{1}}{x_{p}} \\ y_{2} &=&x_{p}\sqrt{\frac{m_{1}m_{2}}{m_{1}^{2}-x_{p}^{4}}}\left[ 1-\frac{% m_{1}}{x_{p}^{2}}-\frac{m_{1}^{2}-x_{p}^{4}}{m_{1}x_{p}^{2}}\right] \end{eqnarray} \end{mathletters} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{F5VDNY0Y.eps}} \caption{For planets internal to the star's Einstein ring, the planetary caustics move on this curve.} \label{Fig caustics position} \end{figure} The first of these is a well known formula (Griest \& Safizadeh, 1997). The second completes the information given from the first and allows a complete individuation of the two caustics. Fig. \ref{Fig caustics position} is a plot of the position of the two caustics as a function of the distance of the planet from the star. When $x_{p}\rightarrow 0$ the caustics approximately move on the lines: \begin{equation} y_{2}=\pm 2\sqrt{\frac{m_{2}}{m_{1}}}y_{1} \end{equation} The two planetary caustics delimitate a region of high de - amplification which can appear in microlensing light curves as negative peaks. The positions of the caustics can give a measure of the size of this region and consequently the size of these negative peaks. \subsection{Cusps} This subsection concludes the study of the perturbative caustics with the analysis of the position of cusps in these caustics. The position of cusps can be important in several studies such as microlensing itself. In fact cusps are surrounded by a region with an amplification even higher than that of fold singularities. They also define the extension and the shape of the caustic. Cusps are defined as the points where the tangent vector of the caustic vanishes. In order to find them we must set \begin{equation} \left\{ \begin{array}{c} y_{1}^{\prime }\left( \vartheta \right) =0 \\ y_{2}^{\prime }\left( \vartheta \right) =0 \end{array} \right. \label{General cusp equation} \end{equation} and resolve this system of equations for $\vartheta $. Let's start with the central caustic. Eqs. (\ref{General cusp equation}) after several steps become: \begin{equation} \left\{ \begin{array}{c} \cos \vartheta \left(\frac{\partial \varepsilon }{\partial \vartheta }- \sum\limits_{i=2}^{n}m_{i}\frac{\left( \Delta _{i2}^{0}\cos \vartheta -\Delta _{i1}^{0}\sin \vartheta \right) \left( \Delta _{i1}^{0}\cos \vartheta +\Delta _{i2}^{0}\sin \vartheta \right) }{\left[ \left( \Delta _{i1}^{0}\right) ^{2}+\left( \Delta _{i2}^{0}\right) ^{2}% \right] ^{2}}\right)=0 \\ \sin \vartheta \left(\frac{\partial \varepsilon }{\partial \vartheta }- \sum\limits_{i=2}^{n}m_{i}\frac{\left( \Delta _{i2}^{0}\cos \vartheta -\Delta _{i1}^{0}\sin \vartheta \right) \left( \Delta _{i1}^{0}\cos \vartheta +\Delta _{i2}^{0}\sin \vartheta \right) }{\left[ \left( \Delta _{i1}^{0}\right) ^{2}+\left( \Delta _{i2}^{0}\right) ^{2}% \right] ^{2}}\right)=0 \end{array} \right. \end{equation} These can simultaneously vanish only if \begin{equation} \frac{\partial \varepsilon }{\partial \vartheta }-\sum\limits_{i=2}^{n}m_{i}% \frac{\left( \Delta _{i2}^{0}\cos \vartheta -\Delta _{i1}^{0}\sin \vartheta \right) \left( \Delta _{i1}^{0}\cos \vartheta +\Delta _{i2}^{0}\sin \vartheta \right) }{\left[ \left( \Delta _{i1}^{0}\right) ^{2}+\left( \Delta _{i2}^{0}\right) ^{2}\right] ^{2}}=0 \end{equation} Explicitly, this equation is: \begin{multline} \sum\limits_{i=2}^{n}\frac{m_{i}}{\left[ \left( \Delta _{i1}^{0}\right) ^{2}+\left( \Delta _{i2}^{0}\right) ^{2}\right] ^{3}}\left\{ 3\left[ \left( \Delta _{i2}^{0}\right) ^{4}-\left( \Delta _{i1}^{0}\right) ^{4}\right] \sin \vartheta \cos \vartheta +\right. \\ +3\Delta _{i1}^{0}\Delta _{i2}^{0}\left[ \left( \Delta _{i1}^{0}\right) ^{2}+\left( \Delta _{i2}^{0}\right) ^{2}\right] \left( \cos ^{2}\vartheta -\sin ^{2}\vartheta \right) + \\ -\Delta _{i1}^{0}\sqrt{m_{1}}\left[ \left( \Delta _{i1}^{0}\right) ^{2}-3\left( \Delta _{i2}^{0}\right) ^{2}\right] \left( 4\sin ^{3}\vartheta -3\sin \vartheta \right) + \\ \left. +\Delta _{i2}^{0}\sqrt{m_{1}}\left[ \left( \Delta _{i2}^{0}\right) ^{2}-3\left( \Delta _{i1}^{0}\right) ^{2}\right] \left( 4\cos ^{3}\vartheta -3\cos \vartheta \right) \right\} \end{multline} which, in despite of its cumbersome aspect, can be exactly solved in the case of the single planet where it yields the four solutions: \begin{multline} \vartheta =0\RIfM@\expandafter\text@\else\expandafter\mbox\fi{; \ }\vartheta =\pi \RIfM@\expandafter\text@\else\expandafter\mbox\fi{; \ } \\ \vartheta =\pm \arccos % \left[ \frac{3m_{1}+3x_{p}^{2}-\sqrt{9m_{1}^{2}-14m_{1}x_{p}^{2}+9x_{p}^{4}}% }{4\sqrt{m_{1}}x_{p}}\right] \end{multline} For planetary caustics, we can proceed in a similar way. Multiplying (\ref {General cusp equation}a) by $\sin \vartheta $, (\ref{General cusp equation}% b) by $\cos \vartheta $ and subtracting, we have: \begin{equation} \frac{m_{1}}{x_{p}^{2}}r\left( \cos ^{2}\vartheta -\sin ^{2}\vartheta \right) +2\frac{m_{1}}{x_{p}^{2}}\frac{\partial r}{\partial \vartheta }\sin \vartheta \cos \vartheta -r+\frac{m_{2}}{r}=0 \end{equation} Multiplying by $r$, we get an equation in $r^{2}$ which is easier to handle: \begin{equation} \frac{m_{1}}{x_{p}^{2}}r^{2}\left( \cos ^{2}\vartheta -\sin ^{2}\vartheta \right) +\frac{m_{1}}{x_{p}^{2}}\frac{\partial r^{2}}{\partial \vartheta }% \sin \vartheta \cos \vartheta -r^{2}+m_{2}=0 \end{equation} Inserting (\ref{Planetary critical curve}) and solving, we have: \begin{equation} \vartheta =0\RIfM@\expandafter\text@\else\expandafter\mbox\fi{; \ }\vartheta =\frac{\pi }{2}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{; \ }\vartheta =\pi \RIfM@\expandafter\text@\else\expandafter\mbox\fi{; \ }\vartheta =\frac{3\pi }{2} \end{equation} on the higher branch, and: \begin{mathletters} \begin{eqnarray} \vartheta &=&\pm \frac{1}{2}\arcsin \left[ \frac{\sqrt{3}x_{p}^{2}}{2m_{1}}% \right] \RIfM@\expandafter\text@\else\expandafter\mbox\fi{;} \\ \vartheta &=&\pi \pm \frac{1}{2}\arcsin \left[ \frac{\sqrt{3% }x_{p}^{2}}{2m_{1}}\right] \RIfM@\expandafter\text@\else\expandafter\mbox\fi{;} \\ \vartheta &=&\frac{\pi }{2}\pm \frac{1}{2}\arcsin \left[ \frac{\sqrt{3}% x_{p}^{2}}{2m_{1}}\right] \RIfM@\expandafter\text@\else\expandafter\mbox\fi{;} \label{Lower planetary cusps 1} \\ \vartheta &=&\frac{3\pi }{2}\pm \frac{1}{2% }\arcsin \left[ \frac{\sqrt{3}x_{p}^{2}}{2m_{1}}\right] \label{Lower planetary cusps 2} \end{eqnarray} \end{mathletters} on the lower. For external planets, the higher branch is complete while the lower is absent, so only the four cusps on the higher branch are actually present. For internal planets, the two branches are real only near $\vartheta =\frac{% \pi }{2}$ and $\vartheta =\frac{3\pi }{2}$. So the higher branch has the two cusps at $\vartheta =\frac{\pi }{2}$ and $\vartheta =\frac{3\pi }{2}$, while in the lower one the four cusps (\ref{Lower planetary cusps 1}) and (\ref{Lower planetary cusps 2}) are real and the others are imaginary. Summing up we have six cusps distributed in such a way as to form two triangular caustics. \section{Microlensing} \noindent Among the numerous forms of gravitational lensing, microlensing is surely one of the most relevant since it opens the possibility of probing the galactic structure through a directly gravitational investigation. Microlensing occurs when the images of a given source, produced by a small lens, are too close (typically less than $10^{-3}$ arcsecs) to be separated by our telescopes. As we cannot see but a point image of the source, the only way to notice a lensing effect is through a variation of the total light flux coming from the observed source. For a point lens mass, this variation was found by Paczynski (1986) who first thought of galactic microlensing as a new observable astronomical phenomenon. For a planetary system the anomalies in amplification patterns do not enjoy a full analytical description. Our aim is to use a perturbative approach to solve this problem and find analytical light curves for stars accompanied by their planets. \subsection{Paczynski's curve} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{E5VDNY0Z.eps}} \caption{Typical microlensing curve for a point lens mass.} \label{Fig Paczynski's curve} \end{figure} Before considering the problem of planetary microlensing, it is useful to review the steps to be followed in order to get amplification light curves in the event of a single mass (Schneider, Ehlers \& Falco, 1992). This will help us in fixing the problems to be faced. In this case, the lens equation takes a very simple aspect: \begin{equation} \mathbf{y}=\mathbf{x}-\frac{m_{1}}{\left| \mathbf{x}\right| ^{2}}% \mathbf{x} \label{Single lens equation} \end{equation} The total amplification is found by summing the amplification of all images. So the first step is to find these images, i.e. the lens equation is to be inverted. Here the task is quite easy, because (\ref{Single lens equation}) reduces to a second degree equation whose solutions are: \begin{mathletters} \label{Paczynski's images} \begin{eqnarray} \mathbf{I}^{+} &=&\frac{\mathbf{y}}{2}\left( 1+\frac{\sqrt{% 4m_{1}+\left| \mathbf{y}\right| ^{2}}}{\left| \mathbf{y}\right| }% \right) \\ \mathbf{I}^{-} &=&\frac{\mathbf{y}}{2}\left( 1-\frac{\sqrt{% 4m_{1}+\left| \mathbf{y}\right| ^{2}}}{\left| \mathbf{y}\right| }% \right) \end{eqnarray} \end{mathletters} The next step is to compute the amplifications corresponding to each of these images. According to (\ref{General amplification}), these are: \begin{equation} \mu _{I^{\pm }}=\frac{1}{\left| \det J\left( \mathbf{I}^{\pm }\right) \right| }=\frac{1}{\left| 1-\frac{m_{1}^{2}}{\left| \mathbf{I}^{\pm }\right| ^{4}}\right| } \end{equation} It is interesting to study the properties of the two images to discover their physical essence (Blandford \& Narayan, 1986). The image $\mathbf{I}% ^{+}$ has positive parity; in the limit of vanishing lensing effect $\left( \left| \mathbf{y}\right| ^{2}\gg m_{1}\right) $ $\mathbf{I}^{+}$ tends to $\mathbf{y}$ and its amplification becomes unitary. Thus $\mathbf{I}% ^{+}$ reduces to the usual image of the source in the absence of lensing. In what follows I'll refer to it as the principal image. $\mathbf{I}^{-}$ has negative parity and in the limit of low lensing goes as $m_{1}/\left| \mathbf{y}\right| $, while its amplification is always $\mu _{I^{-}}=\mu _{I^{+}}-1$. I shall call it secondary image as it disappears when the lensing effect is not present. Both images are aligned on the line connecting the source and the lens: the principal image is always external to the Einstein ring, while the secondary one is internal to it. Now we have to sum up the two amplifications to obtain the so - called amplification map: \begin{equation} \mu _{TOT}=\frac{2m_{1}+\left| \mathbf{y}\right| ^{2}}{\left| \mathbf{y% }\right| \sqrt{4m_{1}+\left| \mathbf{y}\right| ^{2}}} \label{Paczynski's amplification} \end{equation} This function tells us the amplification corresponding to any given position of the source relatively to the lensing object. Of course it only depends on the distance because of the symmetry of the lens. The final step is to make the source move along a rectilinear trajectory to obtain the complete light curves corresponding to the passage of a massive lens near the line of sight of the source (obviously it makes no difference who is moving, what counts is only the relative motion). The distance $% \left| \mathbf{y}\right| $ is: \begin{equation} \left| \mathbf{y}\right| =\sqrt{b^{2}+v_{\bot }^{2}t^{2}} \end{equation} where $b$ is the impact parameter (the closest approach distance) and $% v_{\bot }$ is the projection of the relative speed in a plane orthogonal to the line of sight. A typical light curve is shown in fig. \ref{Fig Paczynski's curve}. The height of the maximum is found by substituting the impact parameter in (\ref {Paczynski's amplification}). It becomes infinite as $b$ tends to zero. Real sources have finite extensions implying integration processes smoothing the peak of the curve (Witt \& Mao, 1994). This cut - off becomes evident when $\left| \mathbf{y}\right| $ is comparable to the source radius. \subsection{The problem of planetary microlensing} In principle, the procedure for attaining microlensing light curves for multiple lenses is the same just expounded for a single point lens. First we should invert the lens application, second we have to compute the amplification of all the images, then sum up to have the amplification map and finally introduce the motion of the source relatively to the lens system. But if we write the lens equation for a star with just one planet placed in $\mathbf{x}_{p}=\left( x_{p},0\right) $: \begin{equation} \mathbf{y}=\mathbf{x}-\frac{m_{1}}{\left| \mathbf{x}\right| ^{2}}% \mathbf{x}-\frac{m_{2}}{\left| \mathbf{x}-\mathbf{x}_{p}\right| ^{2}% }\left( \mathbf{x}-\mathbf{x}_{p}\right) \label{One planet lens equation} \end{equation} we at once see that the inversion is not possible since one must surrender at a fifth degree equation which doesn't allow to find the images produced by such a lens. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{F5VDNY10.eps}} \caption{In presence of a planet, a source (here represented by a four-cornered star) placed outside the caustics generates three images (the three little circles in the figure).} \label{Fig planetary images} \end{figure} A glance at the numerical results can indicate us which way is to be taken in the inversion of the lens application (\ref{One planet lens equation}). When the source is outside the caustics, only three images are present (see fig. \ref{Fig planetary images}). One of them is outside all critical curves and approaches the source when the latter is far enough from the lens. This is indeed the principal image. Another image is inside the star's critical curve. It is easy to understand that when the mass of the planet vanishes this image becomes the star's secondary image. The last image is near the planet (inside the planetary critical curve when the planet is external to the star's Einstein ring). I shall refer to this as the planetary image. It is clear that the presence of the planet slightly perturbs the principal and secondary image of the star, so that their position can be found applying perturbation theory to Paczynski's images. The planetary image is completely perturbative, since it is not present in the zeroth order situation in which the planet is absent, and must be treated separately. When the source threads a caustic, two new images are formed with opposite parities whose effects are similar to those of the planetary image. So, the perturbative analysis is likely to be the key to solve the problem of planetary microlensing. In the following two subsections I will use it to discover the images and their amplification. Finally, I will build amplification light curves and compare them with their numerical counterparts. \subsection{Principal and secondary image} Paczynski's images (\ref{Paczynski's images}) are the starting point for our expansion and will be generically indicated by the symbol $\mathbf{I}% ^{\left( 0\right) }$. Let's write the position of the images to the first order in $m_{2}$ as the sum of Paczynski's image and a small perturbation $\mathbf{\Delta I}$: \begin{equation} \mathbf{I}^{\left( 1\right) }=\mathbf{I}^{\left( 0\right) }+\mathbf{% \Delta I} \end{equation} With this position, in the lens equation expanded to the first order in $% m_{2}$% \begin{mathletters} \begin{eqnarray} y_{1} =I_{1}^{\left( 1\right) }-\frac{m_{1}I_{1}^{\left( 1\right) }}{% \left( I_{1}^{\left( 1\right) }\right) ^{2}+\left( I_{2}^{\left( 1\right) }\right) ^{2}}+ \notag \\ -\frac{m_{2}\left( I_{1}^{\left( 0\right) }-x_{p}\right) }{% \left( I_{1}^{\left( 0\right) }-x_{p}\right) ^{2}+\left( I_{2}^{\left( 0\right) }\right) ^{2}} \\ y_{2} =I_{2}^{\left( 1\right) }-\frac{m_{1}I_{2}^{\left( 1\right) }}{% \left( I_{1}^{\left( 1\right) }\right) ^{2}+\left( I_{2}^{\left( 1\right) }\right) ^{2}}+ \notag \\ -\frac{m_{2}I_{2}^{\left( 0\right) }}{\left( I_{1}^{\left( 0\right) }-x_{p}\right) ^{2}+\left( I_{2}^{\left( 0\right) }\right) ^{2}} \end{eqnarray} \end{mathletters} the planetary term no longer contains the perturbation $\mathbf{\Delta I}$% . Putting: \begin{mathletters} \begin{eqnarray} \Delta y_{1}=\frac{m_{2}\left( I_{1}^{\left( 0\right) }-x_{p}\right) }{% \left( I_{1}^{\left( 0\right) }-x_{p}\right) ^{2}+\left( I_{2}^{\left( 0\right) }\right) ^{2}}\RIfM@\expandafter\text@\else\expandafter\mbox\fi{;} \\ \Delta y_{2}=\frac{m_{2}I_{2}^{\left( 0\right) }}{\left( I_{1}^{\left( 0\right) }-x_{p}\right) ^{2}+\left( I_{2}^{\left( 0\right) }\right) ^{2}} \end{eqnarray} \end{mathletters} and bringing these terms to the left members, we re-gain the structure of the Schwarzschild lens equation (\ref{Single lens equation}) in the variable $\mathbf{I}^{\left( 1\right) }$ for the source position $\mathbf{y}+% \mathbf{\Delta y}$. The planetary induced perturbation can be thus read as a shift in the source position. $\mathbf{I}^{\left( 1\right) }$ has the same expression as $\mathbf{I}^{\left( 0\right) }$ evaluated in $% \mathbf{y}+\mathbf{\Delta y}$ instead of $\mathbf{y}$. The perturbation $\mathbf{\Delta I}$ is found by expanding $\mathbf{I}% ^{\left( 1\right) }$ to the first order in $\mathbf{\Delta y}$: \begin{mathletters} \label{Images perturbation} \begin{eqnarray} \Delta I_{1}^{\pm } =\frac{\Delta y_{1}}{2}\left( 1\pm \frac{\sqrt{% 4m_{1}+y_{1}^{2}+y_{2}^{2}}}{\sqrt{y_{1}^{2}+y_{2}^{2}}}\right) \notag \\ \mp \frac{% 2m_{1}y_{1}\left( y_{1}\Delta y_{1}+y_{2}\Delta y_{2}\right) }{\sqrt{% 4m_{1}+y_{1}^{2}+y_{2}^{2}}\sqrt{\left( y_{1}^{2}+y_{2}^{2}\right) ^{3}}} \\ \Delta I_{2}^{\pm } =\frac{\Delta y_{2}}{2}\left( 1\pm \frac{\sqrt{% 4m_{1}+y_{1}^{2}+y_{2}^{2}}}{\sqrt{y_{1}^{2}+y_{2}^{2}}}\right) \notag \\ \mp \frac{% 2m_{1}y_{2}\left( y_{1}\Delta y_{1}+y_{2}\Delta y_{2}\right) }{\sqrt{% 4m_{1}+y_{1}^{2}+y_{2}^{2}}\sqrt{\left( y_{1}^{2}+y_{2}^{2}\right) ^{3}}} \end{eqnarray} \end{mathletters} The upper signs stand for the principal image and the lower for the secondary. The expansion parameter $m_{2}$ appears through $\mathbf{% \Delta y}$. Now the position of the principal and secondary image are known. The most delicate operation is done and the door to the planetary microlensing is open at last. What remains is only mechanic computation without any conceptual difficulties. The amplification of each image is found by the general formula (\ref {General amplification}) expanded to the first order in $m_{2}$ (I drop the zero from $\mathbf{I}^{\left( 0\right) }$ to simplify notation): \begin{multline} \mu _{I}=\frac{1}{\left| 1-\frac{m_{1}^{2}}{\left( I_{1}^{2}+I_{2}^{2}\right) ^{2}}\right| }\left\{ 1-\frac{4m_{1}^{2}}{\left( I_{1}^{2}+I_{2}^{2}\right) ^{3}}\left( I_{1}\Delta I_{1}+I_{2}\Delta I_{2}\right) +\right. \label{Image amplification} \\ +\frac{2m_{1}m_{2}}{\left( 1-\frac{m_{1}^{2}}{\left( I_{1}^{2}+I_{2}^{2}\right) ^{2}}\right) } \\ \left. \left[ \frac{\left( I_{1}^{2}-I_{2}^{2}\right) \left( \left( I_{1}-x_{p}\right) ^{2}-I_{2}^{2}\right) +4I_{1}\left( I_{1}-x_{p}\right) I_{2}^{2}}{\left( I_{1}^{2}+I_{2}^{2}\right) ^{2}\left( \left( I_{1}-x_{p}\right) ^{2}+I_{2}^{2}\right) ^{2}}\right] \right\} \end{multline} This is the sought formula for the amplification of the images. Paczynski's amplification multiplies the main brackets containing the sum of all perturbations following the zero order solution represented by the unity. Two kinds of perturbations can be recognized: the first is caused by the previously found shift in the image positions $\mathbf{\Delta I}$; the second is the consequence of the change of the function $\det J$ produced by the presence of the planetary term in the lens equation. I have dropped the modulus from the main brackets because its content is always positive since the perturbations are smaller than unity (except for the zones where perturbative method is no longer valid). As usual, the validity of perturbation theory is limited to the regions where perturbations are enough small to make sense. So it is necessary a careful examination of the denominators of all perturbative terms. The shift terms present the distance of the zeroth order image from the origin raised to the sixth power. There's no problem for the principal image which is always far beyond the Einstein ring, but this is not true for the secondary image. However the ``failure'' rises in the limit of vanishing lensing where the amplification of this image is so low to be totally masked by the amplification of the principal image. When the amplification of the secondary image begins to become important, the distance from the origin is largely sufficient to eliminate all the problems and have fine perturbations. The shift $\mathbf{\Delta I}$ becomes infinite when the source passes through the origin; so the region very near the origin is the first to avoid. The displacement $\mathbf{\Delta y}$ diverges when the zeroth order image approaches the planet as could easily be foreseen for a first order perturbation theory. As regards the terms coming from the alteration of $\det J$, there's nothing new; the prescriptions are the same as those from the other terms. In sum we are allowed to use these amplification formulae for all source positions being not too near the origin or generating images too close to the planet. This hardly happens when the source is internal to the caustics. We'll see that very reliable results can be obtained up to very short distances from the centres of the caustics. \subsection{Planetary image} As previously announced, in this subsection I shall deal with the third image. The presence of this image is absolutely tied to that of the planet. Anyway, Paczynski's images can still provide a good starting point for our analysis. In fact, if the planet is very far from the star, it too will behave as a single lens. In this case, the planetary image is nothing but the secondary Paczynski's image for a very low mass. In this limit, its distance from the planet is of order $m_{2}$. So, in our perturbative expansion, we are encouraged to search for images with distances from the planet of order $m_{2}$. Let $\mathbf{I}^{p}$ be the position of the planetary image. We have: \begin{equation} \mathbf{I}^{p}=\mathbf{x}_{p}+\mathbf{\Delta I}^{p} \label{Image planetary expansion} \end{equation} with $\mathbf{\Delta I}^{p}$ of order $m_{2}$. Saving only the lowest order, the lens application reads: \begin{mathletters} \begin{eqnarray} y_{1} &=&x_{p}-\frac{m_{1}}{x_{p}}-\frac{m_{2}\Delta I_{1}^{p}}{\left( \Delta I_{1}^{p}\right) ^{2}+\left( \Delta I_{2}^{p}\right) ^{2}} \\ y_{2} &=&-\frac{m_{2}\Delta I_{2}^{p}}{\left( \Delta I_{1}^{p}\right) ^{2}+\left( \Delta I_{2}^{p}\right) ^{2}} \end{eqnarray} \end{mathletters} These equations can easily be solved. The solution is: \begin{eqnarray} I_{1}^{p} &=&x_{p}-\frac{m_{2}\left( y_{1}-y_{p}\right) }{\left( y_{1}-y_{p}\right) ^{2}+y_{2}^{2}} \\ I_{1}^{p} &=&-\frac{m_{2}y_{2}}{\left( y_{1}-y_{p}\right) ^{2}+y_{2}^{2}} \end{eqnarray} where $y_{p}=x_{p}-\frac{m_{1}}{x_{p}}$\ is the zeroth order position of the planetary caustic already rising in former discussions. As ever the amplification is calculated by expanding (\ref{General amplification}). The lowest order result is: \begin{equation} \mu _{I^{p}}=\frac{m_{2}^{2}}{\left[ \left( y_{1}-y_{p}\right) ^{2}+y_{2}^{2}% \right] ^{2}} \end{equation} Notice how this formula is much more simple than other images amplification. The denominators in these expressions vanish when $\mathbf{y}\rightarrow \mathbf{y}_{p}$. Consequently the perturbative method fails when the source is very close to the centre of the planetary caustic. \subsection{Perturbative light curves in planetary microlensing} Once we have found the amplification for each image, in order to obtain the microlensing amplification map we must sum up the components coming from the three images. However, we see that the contribution to the total amplification of the planetary image is of the second order in $m_{2}$. Since we are only considering first order corrections to Paczynski's curve, this contribution is to be ignored. Therefore, from now on, we shall confine ourselves to the principal and secondary images only. One consideration is for the two hidden images coming out when the source crosses a caustic. If the event regards the planetary caustic, the two images can be found by carrying further the expansion (\ref{Image planetary expansion}). The new images arise from higher order solutions and their amplifications will also be of higher orders. So we don't worry about them. On the contrary, if the source crosses the central caustic, the new images appear near the star's Einstein ring, far from any possible starting point for a perturbative expansion. As we are not taking them into account, we cannot expect to obtain good results inside the central caustic. Anyway, central caustic crossing events are very improbable, since the extension of this caustic is $m_{2}^{2}/m_{1}^{2}$ times the star's Einstein ring. Building light curves presents no difficulty. Chosen one source trajectory, it suffices to parameterize $y_{1}$ and $y_{2}$ in the amplification map properly. This is no longer a function of the radial coordinate because there is no more rotational symmetry. To account for the finite size of the source a simple numerical integration of the perturbative amplification map on the source area at each point of the trajectory can be performed. The curves thus obtained can be compared to numerical ones given by ``inverse ray shooting'' algorithm. All the results I show in this paper regard a system constituted by a star with mass $m_{1}=1$ and a Jovian planet ($m_{2}=10^{-3}$). This choice has been made in order to put in better evidence planetary perturbations and to test the perturbative approach in the least favourable situation. Obviously with Earth - like planets things can only go better. \begin{figure*} \resizebox{12cm}{!}{\includegraphics{ligext05.eps}} \hfill \parbox[b]{55mm}{ \caption{Besides a star with unitary mass placed in the origin, here is a Jovian planet ($m_{2}=10^{-3}$) in $\protect% \mathbf{x}_{p}=\left( 1.2;0\right) $. The trajectory of the source shown in (a) has impact parameter 0.5. (b) is a numerical light curve obtained by ``inverse ray shooting'' for a source 43 times greater than the sun. (c) is the perturbative light curve for a point source and (d) is the same curve after a numerical integration over the source extension.} \label{Fig light curve extern 0.5}} \end{figure*} \begin{figure*} \resizebox{12cm}{!}{\includegraphics{ligext02.eps}} \hfill \parbox[b]{55mm}{ \caption{Same lens system and same figure ordering as before. The impact parameter is 0.2.} \label{Fig light curve extern 0.2}} \end{figure*} \begin{figure*} \resizebox{12cm}{!}{\includegraphics{ligext04.eps}} \hfill \parbox[b]{55mm}{ \caption{The lens system is the same again but the impact parameter 0.4 makes the source cross the planetary caaustic.} \label{Fig light curve extern 0.4}} \end{figure*} \begin{figure*} \resizebox{12cm}{!}{\includegraphics{ligint06.eps}} \hfill \parbox[b]{55mm}{ \caption{Now the planet is in $% \protect\mathbf{x}_{p}=\left( 0.8,0\right) $. The three light curves are obtained in the same ways described in fig. \ref{Fig light curve extern 0.5}% . The impact parameter is 0.6.} \label{Fig light curve internal 0.6}} \end{figure*} \begin{figure*} \resizebox{12cm}{!}{\includegraphics{ligint02.eps}} \hfill \parbox[b]{55mm}{ \caption{Same lens system as in the previous figure. The impact parameter is 0.25.} \label{Fig light curve internal 0.25}} \end{figure*} \begin{figure*} \resizebox{12cm}{!}{\includegraphics{ligint04.eps}} \hfill \parbox[b]{55mm}{ \caption{Caustic crossing with impact parameter 0.4.} \label{Fig light curve internal 0.4}} \end{figure*} Let's start with an external planet. In fig. \ref{Fig light curve extern 0.5} the planet is in $x_{p}=1.2$. The trajectory chosen for this first test is shown in fig. \ref{Fig light curve extern 0.5}a and has impact parameter 0.5. The numerically attained light curve is displayed in fig. \ref{Fig light curve extern 0.5}b. The source used for this curve has radius 0.045. In a standard observation towards the bulge of the galaxy ($D_{OL}\sim 8kpc$% , $D_{OS}\sim 10kpc$), this value would correspond to a giant about 43 times greater than the sun. Here the presence of the planet is responsible for the little peak on the left of the maximum of the curve. Fig. \ref{Fig light curve extern 0.5}c\ represents the perturbative light curve for a point source moving along the same trajectory. If we perform the numerical integration of the perturbative amplification map, as previously said, the perturbative light curve \ref{Fig light curve extern 0.5}d becomes indistinguishable from the numerical one. This is a very encouraging result, so let's choose other trajectories to see other tests. In fig. \ref{Fig light curve extern 0.2} the position of the planet is the same but the trajectory passes between the planetary caustic and the central caustic at a minimum distance of 0.2. The peak in the numerical curve \ref{Fig light curve extern 0.2}b is very close to the maximum. The point source perturbative curve \ref{Fig light curve extern 0.2}% c presents a sharp peak which assumes the right proportions after the integration in \ref{Fig light curve extern 0.2}d. At this point, let's see what happens when the source crosses the planetary caustic. In fig. \ref{Fig light curve extern 0.4}a, the impact parameter 0.4 allows the crossing. The peak in the numerical curve \ref{Fig light curve extern 0.4}b becomes considerably high.\ In the point source perturbative curve \ref{Fig light curve extern 0.4}c\ the peak is very sharp (it would diverge at the centre of the caustic $\mathbf{y}_{p}$). However, the integration over the source surface still succeeds in reporting this peak to the right size and shape. Now, let's consider an internal planet\ ($x_{p}=0.8$). The region between the couple of planetary caustic is characteristic for its high de-amplification. This produces negative peaks on light curves such as the one shown in fig. \ref{Fig light curve internal 0.6}b corresponding to the trajectory in fig. \ref{Fig light curve internal 0.6}a. It is interesting to see that the perturbative method reproduces this situation with the same great accuracy proved in the former situations. As ever, the point source peak in \ref{Fig light curve internal 0.6}c is smoothed by finite source effect in \ref{Fig light curve internal 0.6}d. In\ fig. \ref{Fig light curve internal 0.25} the impact parameter is 0.25 and things go perfectly as previously. Finally, let's consider caustic crossing in this case. Fig. \ref{Fig light curve internal 0.4}a shows a trajectory very close to the planetary caustics. The ``inverse ray shooting'' curve \ref{Fig light curve internal 0.4}b presents a large de-amplification preceded and followed by little positive peaks. The perturbative curve \ref{Fig light curve internal 0.4}c is characterized by the same situation but the de-amplification is so high to make the total amplification become (unphysically) negative. Now let's see what happens with a finite source. Because of its extension, part of the source hits the centre of the caustic $\mathbf{y}_{p}$ where the perturbative amplification map diverges. This is a hard problem for the numerical integration which becomes very unstable in this zone, so the bottom of the de-amplification region of the light curve \ref{Fig light curve internal 0.4}d cannot be taken as significant. However, we see that things go fairly well even in this extreme situation. \section{Conclusions} \noindent The success of perturbative theory in planetary lensing cannot but impress for the simplicity of the calculations involved and the surprising accuracy of the results even in the hardest situations. In the derivation of the caustics of a planetary system, by a simple idea and very few passages the complete structure of these curves has been easily obtained. The almost complete insensitivity of the perturbative approach to the number of the planets allows complete descriptions of planetary systems without any loss of generality. Also many important physical assertions can be stated thanks to these results. The fact that the shape of the central caustic is largely given by a linear superposition of the effects of \ the single planets is indeed remarkable. In planetary microlensing the results are even exalting. The perturbative amplification map allows the construction of very fine light curves. In the derivation of the amplification map I have dealt with only one planet for the sake of simplicity. Yet the generalization to an arbitrary number of planets is immediate because in the first order domain a superposition principle is here valid as well. For point sources, light curves can be attained in a completely analytical way, while for finite sources I have resorted to numerical integration until now. Work to englobe finite source effect in the analytical description is in progress. When these curves are available, the extraction of parameters of planetary systems from microlensing light curves will start on more solid analytical bases. Also it could be possible to use the analytical expressions in experimental fits, though the large number of parameters would greatly affect the uncertainties in their determination. \begin{acknowledgements} I would like to thank Salvatore Capozziello, Gaetano Lambiase and Gaetano Scarpetta for their valuable suggestions and interesting discussions on this matter. Also I greatly wish to acknowledge Giovanni Covone for guiding me in my initiation to planetary microlensing. \end{acknowledgements} \section*{Abstract (Not appropriate in this style!)}% \else \small \begin{center}{\bf Abstract\vspace{-.5em}\vspace{\z@}}\end{center}% \quotation \fi }% }{% }% \@ifundefined{endabstract}{\def\endabstract {\if@twocolumn\else\endquotation\fi}}{}% \@ifundefined{maketitle}{\def\maketitle#1{}}{}% \@ifundefined{affiliation}{\def\affiliation#1{}}{}% \@ifundefined{proof}{\def\proof{\noindent{\bfseries Proof. }}}{}% \@ifundefined{endproof}{\def\endproof{\mbox{\ \rule{.1in}{.1in}}}}{}% \@ifundefined{newfield}{\def\newfield#1#2{}}{}% \@ifundefined{chapter}{\def\chapter#1{\par(Chapter head:)#1\par }% \newcount\c@chapter}{}% \@ifundefined{part}{\def\part#1{\par(Part head:)#1\par }}{}% \@ifundefined{section}{\def\section#1{\par(Section head:)#1\par }}{}% \@ifundefined{subsection}{\def\subsection#1% {\par(Subsection head:)#1\par }}{}% \@ifundefined{subsubsection}{\def\subsubsection#1% {\par(Subsubsection head:)#1\par }}{}% \@ifundefined{paragraph}{\def\paragraph#1% {\par(Subsubsubsection head:)#1\par }}{}% \@ifundefined{subparagraph}{\def\subparagraph#1% {\par(Subsubsubsubsection head:)#1\par }}{}% \@ifundefined{therefore}{\def\therefore{}}{}% \@ifundefined{backepsilon}{\def\backepsilon{}}{}% \@ifundefined{yen}{\def\yen{\hbox{\rm\rlap=Y}}}{}% \@ifundefined{registered}{% \def\registered{\relax\ifmmode{}\r@gistered \else$\m@th\r@gistered$\fi}% \def\r@gistered{^{\ooalign {\hfil\raise.07ex\hbox{$\scriptstyle\rm\RIfM@\expandafter\text@\else\expandafter\mbox\fi{R}$}\hfil\crcr \mathhexbox20D}}}}{}% \@ifundefined{Eth}{\def\Eth{}}{}% \@ifundefined{eth}{\def\eth{}}{}% \@ifundefined{Thorn}{\def\Thorn{}}{}% \@ifundefined{thorn}{\def\thorn{}}{}% \def\TEXTsymbol#1{\mbox{$#1$}}% \@ifundefined{degree}{\def\degree{{}^{\circ}}}{}% \newdimen\theight \def\Column{% \vadjust{\setbox\z@=\hbox{\scriptsize\quad\quad tcol}% \theight=\ht\z@\advance\theight by \dp\z@\advance\theight by \lineskip \kern -\theight \vbox to \theight{% \rightline{\rlap{\box\z@}}% \vss }% }% }% \def\qed{% \ifhmode\unskip\nobreak\fi\ifmmode\ifinner\else\hskip5\p@\fi\fi \hbox{\hskip5\p@\vrule width4\p@ height6\p@ depth1.5\p@\hskip\p@}% }% \def\cents{\hbox{\rm\rlap/c}}% \def\miss{\hbox{\vrule height2\p@ width 2\p@ depth\z@}}% \def\vvert{\Vert \def\tcol#1{{\baselineskip=6\p@ \vcenter{#1}} \Column} % \def\dB{\hbox{{}} \def\mB#1{\hbox{$#1$} \def\nB#1{\hbox{#1} \@ifundefined{note}{\def\note{$^{\dag}}}{}% \defLaTeX2e{LaTeX2e} \ifx\fmtnameLaTeX2e \DeclareOldFontCommand{\rm}{\normalfont\rmfamily}{\mathrm} \DeclareOldFontCommand{\sf}{\normalfont\sffamily}{\mathsf} \DeclareOldFontCommand{\tt}{\normalfont\ttfamily}{\mathtt} \DeclareOldFontCommand{\bf}{\normalfont\bfseries}{\mathbf} \DeclareOldFontCommand{\it}{\normalfont\itshape}{\mathit} \DeclareOldFontCommand{\sl}{\normalfont\slshape}{\@nomath\sl} \DeclareOldFontCommand{\sc}{\normalfont\scshape}{\@nomath\sc} \fi \def\alpha{{\Greekmath 010B}}% \def\beta{{\Greekmath 010C}}% \def\gamma{{\Greekmath 010D}}% \def\delta{{\Greekmath 010E}}% \def\epsilon{{\Greekmath 010F}}% \def\zeta{{\Greekmath 0110}}% \def\eta{{\Greekmath 0111}}% \def\theta{{\Greekmath 0112}}% \def\iota{{\Greekmath 0113}}% \def\kappa{{\Greekmath 0114}}% \def\lambda{{\Greekmath 0115}}% \def\mu{{\Greekmath 0116}}% \def\nu{{\Greekmath 0117}}% \def\xi{{\Greekmath 0118}}% \def\pi{{\Greekmath 0119}}% \def\rho{{\Greekmath 011A}}% \def\sigma{{\Greekmath 011B}}% \def\tau{{\Greekmath 011C}}% \def\upsilon{{\Greekmath 011D}}% \def\phi{{\Greekmath 011E}}% \def\chi{{\Greekmath 011F}}% \def\psi{{\Greekmath 0120}}% \def\omega{{\Greekmath 0121}}% \def\varepsilon{{\Greekmath 0122}}% \def\vartheta{{\Greekmath 0123}}% \def\varpi{{\Greekmath 0124}}% \def\varrho{{\Greekmath 0125}}% \def\varsigma{{\Greekmath 0126}}% \def\varphi{{\Greekmath 0127}}% \def{\Greekmath 0272}{{\Greekmath 0272}} \def\FindBoldGroup{% {\setbox0=\hbox{$\mathbf{x\global\edef\theboldgroup{\the\mathgroup}}$}}% } \def\Greekmath#1#2#3#4{% \if@compatibility \ifnum\mathgroup=\symbold \mathchoice{\mbox{\boldmath$\displaystyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\textstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptscriptstyle\mathchar"#1#2#3#4$}}% \else \mathchar"#1#2#3# \fi \else \FindBoldGroup \ifnum\mathgroup=\theboldgroup \mathchoice{\mbox{\boldmath$\displaystyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\textstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptstyle\mathchar"#1#2#3#4$}}% {\mbox{\boldmath$\scriptscriptstyle\mathchar"#1#2#3#4$}}% \else \mathchar"#1#2#3# \fi \fi} \newif\ifGreekBold \GreekBoldfalse \let\SAVEPBF=\pbf \def\pbf{\GreekBoldtrue\SAVEPBF}% \@ifundefined{mathletters}{% \newcounter{equationnumber} \def\mathletters{% \addtocounter{equation}{1} \edef\@currentlabel{\arabic{equation}}% \setcounter{equationnumber}{\c@equation} \setcounter{equation}{0}% \edef\arabic{equation}{\@currentlabel\noexpand\alph{equation}}% } \def\endmathletters{% \setcounter{equation}{\value{equationnumber}}% } }{} \@ifundefined{BibTeX}{% \def\BibTeX{{\rm B\kern-.05em{\sc i\kern-.025em b}\kern-.08em T\kern-.1667em\lower.7ex\hbox{E}\kern-.125emX}}}{}% \@ifundefined{AmS}% {\def\AmS{{\protect\usefont{OMS}{cmsy}{m}{n}% A\kern-.1667em\lower.5ex\hbox{M}\kern-.125emS}}}{}% \@ifundefined{AmSTeX}{\def\AmSTeX{\protect\AmS-\protect\TeX\@}}{}% \def\@@eqncr{\let\@tempa\relax \ifcase\@eqcnt \def\@tempa{& & &}\or \def\@tempa{& &}% \else \def\@tempa{&}\fi \@tempa \if@eqnsw \iftag@ \@taggnum \else \@eqnnum\stepcounter{equation}% \fi \fi \global\@ifnextchar*{\@tagstar}{\@tag}@false \global\@eqnswtrue \global\@eqcnt\z@\cr} \def\@ifnextchar*{\@TCItagstar}{\@TCItag}{\@ifnextchar*{\@TCItagstar}{\@TCItag}} \def\@TCItag#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{(#1)}} \def\@TCItagstar*#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{#1}} \def\tfrac#1#2{{\textstyle {#1 \over #2}}}% \def\dfrac#1#2{{\displaystyle {#1 \over #2}}}% \def\binom#1#2{{#1 \choose #2}}% \def\tbinom#1#2{{\textstyle {#1 \choose #2}}}% \def\dbinom#1#2{{\displaystyle {#1 \choose #2}}}% \def\QATOP#1#2{{#1 \atop #2}}% \def\QTATOP#1#2{{\textstyle {#1 \atop #2}}}% \def\QDATOP#1#2{{\displaystyle {#1 \atop #2}}}% \def\QABOVE#1#2#3{{#2 \above#1 #3}}% \def\QTABOVE#1#2#3{{\textstyle {#2 \above#1 #3}}}% \def\QDABOVE#1#2#3{{\displaystyle {#2 \above#1 #3}}}% \def\QOVERD#1#2#3#4{{#3 \overwithdelims#1#2 #4}}% \def\QTOVERD#1#2#3#4{{\textstyle {#3 \overwithdelims#1#2 #4}}}% \def\QDOVERD#1#2#3#4{{\displaystyle {#3 \overwithdelims#1#2 #4}}}% \def\QATOPD#1#2#3#4{{#3 \atopwithdelims#1#2 #4}}% \def\QTATOPD#1#2#3#4{{\textstyle {#3 \atopwithdelims#1#2 #4}}}% \def\QDATOPD#1#2#3#4{{\displaystyle {#3 \atopwithdelims#1#2 #4}}}% \def\QABOVED#1#2#3#4#5{{#4 \abovewithdelims#1#2#3 #5}}% \def\QTABOVED#1#2#3#4#5{{\textstyle {#4 \abovewithdelims#1#2#3 #5}}}% \def\QDABOVED#1#2#3#4#5{{\displaystyle {#4 \abovewithdelims#1#2#3 #5}}}% \def\tint{\mathop{\textstyle \int}}% \def\tiint{\mathop{\textstyle \iint }}% \def\tiiint{\mathop{\textstyle \iiint }}% \def\tiiiint{\mathop{\textstyle \iiiint }}% \def\tidotsint{\mathop{\textstyle \idotsint }}% \def\toint{\mathop{\textstyle \oint}}% \def\tsum{\mathop{\textstyle \sum }}% \def\tprod{\mathop{\textstyle \prod }}% \def\tbigcap{\mathop{\textstyle \bigcap }}% \def\tbigwedge{\mathop{\textstyle \bigwedge }}% \def\tbigoplus{\mathop{\textstyle \bigoplus }}% \def\tbigodot{\mathop{\textstyle \bigodot }}% \def\tbigsqcup{\mathop{\textstyle \bigsqcup }}% \def\tcoprod{\mathop{\textstyle \coprod }}% \def\tbigcup{\mathop{\textstyle \bigcup }}% \def\tbigvee{\mathop{\textstyle \bigvee }}% \def\tbigotimes{\mathop{\textstyle \bigotimes }}% \def\tbiguplus{\mathop{\textstyle \biguplus }}% \def\dint{\mathop{\displaystyle \int}}% \def\diint{\mathop{\displaystyle \iint }}% \def\diiint{\mathop{\displaystyle \iiint }}% \def\diiiint{\mathop{\displaystyle \iiiint }}% \def\didotsint{\mathop{\displaystyle \idotsint }}% \def\doint{\mathop{\displaystyle \oint}}% \def\dsum{\mathop{\displaystyle \sum }}% \def\dprod{\mathop{\displaystyle \prod }}% \def\dbigcap{\mathop{\displaystyle \bigcap }}% \def\dbigwedge{\mathop{\displaystyle \bigwedge }}% \def\dbigoplus{\mathop{\displaystyle \bigoplus }}% \def\dbigodot{\mathop{\displaystyle \bigodot }}% \def\dbigsqcup{\mathop{\displaystyle \bigsqcup }}% \def\dcoprod{\mathop{\displaystyle \coprod }}% \def\dbigcup{\mathop{\displaystyle \bigcup }}% \def\dbigvee{\mathop{\displaystyle \bigvee }}% \def\dbigotimes{\mathop{\displaystyle \bigotimes }}% \def\dbiguplus{\mathop{\displaystyle \biguplus }}% \ifx\ds@amstex\relax \message{amstex already loaded}\makeatother\endinpu \else \@ifpackageloaded{amsmath}% {\message{amsmath already loaded}\makeatother\endinput} {} \@ifpackageloaded{amstex}% {\message{amstex already loaded}\makeatother\endinput} {} \@ifpackageloaded{amsgen}% {\message{amsgen already loaded}\makeatother\endinput} {} \fi \let\DOTSI\relax \def\RIfM@{\relax\ifmmode}% \def\FN@{\futurelet\next}% \newcount\intno@ \def\iint{\DOTSI\intno@\tw@\FN@\ints@}% \def\iiint{\DOTSI\intno@\thr@@\FN@\ints@}% \def\iiiint{\DOTSI\intno@4 \FN@\ints@}% \def\idotsint{\DOTSI\intno@\z@\FN@\ints@}% \def\ints@{\findlimits@\ints@@}% \newif\iflimtoken@ \newif\iflimits@ \def\findlimits@{\limtoken@true\ifx\next\limits\limits@true \else\ifx\next\nolimits\limits@false\else \limtoken@false\ifx\ilimits@\nolimits\limits@false\else \ifinner\limits@false\else\limits@true\fi\fi\fi\fi}% \def\multint@{\int\ifnum\intno@=\z@\intdots@ \else\intkern@\fi \ifnum\intno@>\tw@\int\intkern@\fi \ifnum\intno@>\thr@@\int\intkern@\fi \int \def\multintlimits@{\intop\ifnum\intno@=\z@\intdots@\else\intkern@\fi \ifnum\intno@>\tw@\intop\intkern@\fi \ifnum\intno@>\thr@@\intop\intkern@\fi\intop}% \def\intic@{% \mathchoice{\hskip.5em}{\hskip.4em}{\hskip.4em}{\hskip.4em}}% \def\negintic@{\mathchoice {\hskip-.5em}{\hskip-.4em}{\hskip-.4em}{\hskip-.4em}}% \def\ints@@{\iflimtoken@ \def\ints@@@{\iflimits@\negintic@ \mathop{\intic@\multintlimits@}\limits \else\multint@\nolimits\fi \eat@ \else \def\ints@@@{\iflimits@\negintic@ \mathop{\intic@\multintlimits@}\limits\else \multint@\nolimits\fi}\fi\ints@@@}% \def\intkern@{\mathchoice{\!\!\!}{\!\!}{\!\!}{\!\!}}% \def\plaincdots@{\mathinner{\cdotp\cdotp\cdotp}}% \def\intdots@{\mathchoice{\plaincdots@}% {{\cdotp}\mkern1.5mu{\cdotp}\mkern1.5mu{\cdotp}}% {{\cdotp}\mkern1mu{\cdotp}\mkern1mu{\cdotp}}% {{\cdotp}\mkern1mu{\cdotp}\mkern1mu{\cdotp}}}% \def\RIfM@{\relax\protect\ifmmode} \def\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\RIfM@\expandafter\RIfM@\expandafter\text@\else\expandafter\mbox\fi@\else\expandafter\mbox\fi} \let\nfss@text\RIfM@\expandafter\text@\else\expandafter\mbox\fi \def\RIfM@\expandafter\text@\else\expandafter\mbox\fi@#1{\mathchoice {\textdef@\displaystyle\f@size{#1}}% {\textdef@\textstyle\tf@size{\firstchoice@false #1}}% {\textdef@\textstyle\sf@size{\firstchoice@false #1}}% {\textdef@\textstyle \ssf@size{\firstchoice@false #1}}% \glb@settings} \def\textdef@#1#2#3{\hbox{{% \everymath{#1}% \let\f@size#2\selectfont #3}}} \newif\iffirstchoice@ \firstchoice@true \def\Let@{\relax\iffalse{\fi\let\\=\cr\iffalse}\fi}% \def\vspace@{\def\vspace##1{\crcr\noalign{\vskip##1\relax}}}% \def\multilimits@{\bgroup\vspace@\Let@ \baselineskip\fontdimen10 \scriptfont\tw@ \advance\baselineskip\fontdimen12 \scriptfont\tw@ \lineskip\thr@@\fontdimen8 \scriptfont\thr@@ \lineskiplimit\lineskip \vbox\bgroup\ialign\bgroup\hfil$\m@th\scriptstyle{##}$\hfil\crcr}% \def\Sb{_\multilimits@}% \def\endSb{\crcr\egroup\egroup\egroup}% \def\Sp{^\multilimits@}% \let\endSp\endSb \newdimen\ex@ \ex@.2326ex \def\rightarrowfill@#1{$#1\m@th\mathord-\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill \mkern-6mu\mathord\rightarrow$}% \def\leftarrowfill@#1{$#1\m@th\mathord\leftarrow\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill\mkern-6mu\mathord-$}% \def\leftrightarrowfill@#1{$#1\m@th\mathord\leftarrow \mkern-6mu\cleaders \hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill \mkern-6mu\mathord\rightarrow$}% \def\overrightarrow{\mathpalette\overrightarrow@}% \def\overrightarrow@#1#2{\vbox{\ialign{##\crcr\rightarrowfill@#1\crcr \noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}% \let\overarrow\overrightarrow \def\overleftarrow{\mathpalette\overleftarrow@}% \def\overleftarrow@#1#2{\vbox{\ialign{##\crcr\leftarrowfill@#1\crcr \noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}% \def\overleftrightarrow{\mathpalette\overleftrightarrow@}% \def\overleftrightarrow@#1#2{\vbox{\ialign{##\crcr \leftrightarrowfill@#1\crcr \noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}% \def\underrightarrow{\mathpalette\underrightarrow@}% \def\underrightarrow@#1#2{\vtop{\ialign{##\crcr$\m@th\hfil#1#2\hfil $\crcr\noalign{\nointerlineskip}\rightarrowfill@#1\crcr}}}% \let\underarrow\underrightarrow \def\underleftarrow{\mathpalette\underleftarrow@}% \def\underleftarrow@#1#2{\vtop{\ialign{##\crcr$\m@th\hfil#1#2\hfil $\crcr\noalign{\nointerlineskip}\leftarrowfill@#1\crcr}}}% \def\underleftrightarrow{\mathpalette\underleftrightarrow@}% \def\underleftrightarrow@#1#2{\vtop{\ialign{##\crcr$\m@th \hfil#1#2\hfil$\crcr \noalign{\nointerlineskip}\leftrightarrowfill@#1\crcr}}}% \def\qopnamewl@#1{\mathop{\operator@font#1}\nlimits@} \let\nlimits@\displaylimits \def\setboxz@h{\setbox\z@\hbox} \def\varlim@#1#2{\mathop{\vtop{\ialign{##\crcr \hfil$#1\m@th\operator@font lim$\hfil\crcr \noalign{\nointerlineskip}#2#1\crcr \noalign{\nointerlineskip\kern-\ex@}\crcr}}}} \def\rightarrowfill@#1{\m@th\setboxz@h{$#1-$}\ht\z@\z@ $#1\copy\z@\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\box\z@\mkern-2mu$}\hfill \mkern-6mu\mathord\rightarrow$} \def\leftarrowfill@#1{\m@th\setboxz@h{$#1-$}\ht\z@\z@ $#1\mathord\leftarrow\mkern-6mu\cleaders \hbox{$#1\mkern-2mu\copy\z@\mkern-2mu$}\hfill \mkern-6mu\box\z@$} \def\qopnamewl@{proj\,lim}{\qopnamewl@{proj\,lim}} \def\qopnamewl@{inj\,lim}{\qopnamewl@{inj\,lim}} \def\mathpalette\varlim@\rightarrowfill@{\mathpalette\varlim@\rightarrowfill@} \def\mathpalette\varlim@\leftarrowfill@{\mathpalette\varlim@\leftarrowfill@} \def\mathpalette\varliminf@{}{\mathpalette\mathpalette\varliminf@{}@{}} \def\mathpalette\varliminf@{}@#1{\mathop{\underline{\vrule\@depth.2\ex@\@width\z@ \hbox{$#1\m@th\operator@font lim$}}}} \def\mathpalette\varlimsup@{}{\mathpalette\mathpalette\varlimsup@{}@{}} \def\mathpalette\varlimsup@{}@#1{\mathop{\overline {\hbox{$#1\m@th\operator@font lim$}}}} \def\stackunder#1#2{\mathrel{\mathop{#2}\limits_{#1}}}% \begingroup \catcode `|=0 \catcode `[= 1 \catcode`]=2 \catcode `\{=12 \catcode `\}=12 \catcode`\\=12 |gdef|@alignverbatim#1\end{align}[#1|end[align]] |gdef|@salignverbatim#1\end{align*}[#1|end[align*]] |gdef|@alignatverbatim#1\end{alignat}[#1|end[alignat]] |gdef|@salignatverbatim#1\end{alignat*}[#1|end[alignat*]] |gdef|@xalignatverbatim#1\end{xalignat}[#1|end[xalignat]] |gdef|@sxalignatverbatim#1\end{xalignat*}[#1|end[xalignat*]] |gdef|@gatherverbatim#1\end{gather}[#1|end[gather]] |gdef|@sgatherverbatim#1\end{gather*}[#1|end[gather*]] |gdef|@gatherverbatim#1\end{gather}[#1|end[gather]] |gdef|@sgatherverbatim#1\end{gather*}[#1|end[gather*]] |gdef|@multilineverbatim#1\end{multiline}[#1|end[multiline]] |gdef|@smultilineverbatim#1\end{multiline*}[#1|end[multiline*]] |gdef|@arraxverbatim#1\end{arrax}[#1|end[arrax]] |gdef|@sarraxverbatim#1\end{arrax*}[#1|end[arrax*]] |gdef|@tabulaxverbatim#1\end{tabulax}[#1|end[tabulax]] |gdef|@stabulaxverbatim#1\end{tabulax*}[#1|end[tabulax*]] |endgroup \def\align{\@verbatim \frenchspacing\@vobeyspaces \@alignverbatim You are using the "align" environment in a style in which it is not defined.} \let\endalign=\endtrivlist \@namedef{align*}{\@verbatim\@salignverbatim You are using the "align*" environment in a style in which it is not defined.} \expandafter\let\csname endalign*\endcsname =\endtrivlist \def\alignat{\@verbatim \frenchspacing\@vobeyspaces \@alignatverbatim You are using the "alignat" environment in a style in which it is not defined.} \let\endalignat=\endtrivlist \@namedef{alignat*}{\@verbatim\@salignatverbatim You are using the "alignat*" environment in a style in which it is not defined.} \expandafter\let\csname endalignat*\endcsname =\endtrivlist \def\xalignat{\@verbatim \frenchspacing\@vobeyspaces \@xalignatverbatim You are using the "xalignat" environment in a style in which it is not defined.} \let\endxalignat=\endtrivlist \@namedef{xalignat*}{\@verbatim\@sxalignatverbatim You are using the "xalignat*" environment in a style in which it is not defined.} \expandafter\let\csname endxalignat*\endcsname =\endtrivlist \def\gather{\@verbatim \frenchspacing\@vobeyspaces \@gatherverbatim You are using the "gather" environment in a style in which it is not defined.} \let\endgather=\endtrivlist \@namedef{gather*}{\@verbatim\@sgatherverbatim You are using the "gather*" environment in a style in which it is not defined.} \expandafter\let\csname endgather*\endcsname =\endtrivlist \def\multiline{\@verbatim \frenchspacing\@vobeyspaces \@multilineverbatim You are using the "multiline" environment in a style in which it is not defined.} \let\endmultiline=\endtrivlist \@namedef{multiline*}{\@verbatim\@smultilineverbatim You are using the "multiline*" environment in a style in which it is not defined.} \expandafter\let\csname endmultiline*\endcsname =\endtrivlist \def\arrax{\@verbatim \frenchspacing\@vobeyspaces \@arraxverbatim You are using a type of "array" construct that is only allowed in AmS-LaTeX.} \let\endarrax=\endtrivlist \def\tabulax{\@verbatim \frenchspacing\@vobeyspaces \@tabulaxverbatim You are using a type of "tabular" construct that is only allowed in AmS-LaTeX.} \let\endtabulax=\endtrivlist \@namedef{arrax*}{\@verbatim\@sarraxverbatim You are using a type of "array*" construct that is only allowed in AmS-LaTeX.} \expandafter\let\csname endarrax*\endcsname =\endtrivlist \@namedef{tabulax*}{\@verbatim\@stabulaxverbatim You are using a type of "tabular*" construct that is only allowed in AmS-LaTeX.} \expandafter\let\csname endtabulax*\endcsname =\endtrivlist \def\endequation{% \ifmmode\ifinner \iftag@ \addtocounter{equation}{-1} $\hfil \displaywidth\linewidth\@taggnum\egroup \endtrivlist \global\@ifnextchar*{\@tagstar}{\@tag}@false \global\@ignoretrue \else $\hfil \displaywidth\linewidth\@eqnnum\egroup \endtrivlist \global\@ifnextchar*{\@tagstar}{\@tag}@false \global\@ignoretrue \fi \else \iftag@ \addtocounter{equation}{-1} \eqno \hbox{\@taggnum} \global\@ifnextchar*{\@tagstar}{\@tag}@false% $$\global\@ignoretrue \else \eqno \hbox{\@eqnnum $$\global\@ignoretrue \fi \fi\fi } \newif\iftag@ \@ifnextchar*{\@tagstar}{\@tag}@false \def\@ifnextchar*{\@TCItagstar}{\@TCItag}{\@ifnextchar*{\@TCItagstar}{\@TCItag}} \def\@TCItag#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{(#1)}} \def\@TCItagstar*#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{#1}} \@ifundefined{tag}{ \def\@ifnextchar*{\@tagstar}{\@tag}{\@ifnextchar*{\@tagstar}{\@tag}} \def\@tag#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{(#1)}} \def\@tagstar*#1{% \global\@ifnextchar*{\@tagstar}{\@tag}@true \global\def\@taggnum{#1}} }{} \makeatother \endinput
\section{Introduction} \label{sec-intro} Search for "new physics" has always been one of the most exciting subjects in the field of particle physics. The results presented in 1997 by the H1 \cite{h1_97} and ZEUS \cite{zeus97} experiments at HERA electrified the physics community. Both experiments reported an excess of events in positron-proton Neutral Current Deep Inelastic Scattering (NC DIS) at very high momentum transfer scales $Q^{2}$, as compared with the predictions of the Standard Model. Unfortunately, in spite of the significant increase in the integrated data luminosity, these results have not been confirmed nor contradicted\cite{h1_98,zeus98}. The effect can be just due to a statistical fluctuation, but can also be a first sign of some "new physics". In 1998 HERA experiments started again\footnote{Previously HERA run in electron-proton mode in 1992-94.} to collect electron-proton data aiming for integrated luminosity comparable with that of the positron-proton data. The first results are expected soon. The aim of the presented analysis is to review experimental and theoretical constraints on possible signals of "new physics" at HERA and extract limits on a new effects to be seen in the new HERA $e^-p$ data. Limits corresponding to other present and future high-energy experiments are also considered. The contact interaction models, used as the general framework for this analysis are described in section \ref{sec-model}. In section \ref{sec-data} the relevant data from HERA, LEP, the Tevatron and other experiments are briefly described. Methods used to compare data with contact interaction model predictions are discussed in section \ref{sec-method}. The results of analysis within different contact interaction models, including extracted limits on the mass scale of new interactions, are presented in section \ref{sec-results}. Predictions for the future discovery potential at HERA, as well as at LEP and the Tevatron are discussed in section~\ref{sec-predictions}. The analysis presented here is based on the approach suggested in \cite{ciglob}. Significant work has been done to improve the treatment of experimental data, including a proper interpretation of statistical and systematic errors as well as acceptance cuts and smearing. \section{Contact Interactions} \label{sec-model} Four-fermion contact interactions are an effective theory, which allows us to describe, in the most general way, possible low energy effects coming from "new physics" at much higher energy scales. This includes the possible existence of second generation heavy weak bosons, leptoquarks as well as electron and quark compositeness \cite{cidef,cihera}. Contact interactions can be represented as additional terms in the Standard Model Lagrangian~\cite{cihera}: \begin{eqnarray} L_{CI} & = & \eta_{s} (\bar{e}_{L} e_{R} )(\bar{q}_{L} q_{R}) + \eta_{s}' (\bar{e}_{L} e_{R} )(\bar{q}_{R} q_{L}) + h.c. \hspace{2.5cm} \rm scalar \label{eq-lagr}\\[3mm] & + & \sum_{i,j=L,R} \eta_{ij} (\bar{e}_{i} \gamma^{\mu} e_{i} ) (\bar{q}_{j} \gamma_{\mu} q_{j}) \hspace{5.1cm} \rm vector \nonumber \\[2mm] & + & \eta_{T} (\bar{e}_{L} \sigma^{\mu \nu} e_{R} ) (\bar{q}_{L} \sigma_{\mu \nu} q_{R} ) + h.c. \hspace{4.4cm} \rm tensor \nonumber \end{eqnarray} where subsequent lines describe the scalar, vector and tensor contact interaction terms respectively. As very strong limits have been already placed on both scalar and tensor terms \cite{cihera} this paper considers vector terms only. The influence of the vector contact interactions on the $ep$ NC DIS cross-section can be described as an additional term in the tree level $eq \rightarrow eq$ scattering amplitude~\cite{ciglob}: \begin{eqnarray} M^{e_i q_j \rightarrow e_i q_j}(t) & = & - \frac{4 \pi \alpha_{em} e_{q}}{t} \; + \; \frac{4 \pi \alpha_{em}}{sin^{2}\theta_{W} \cdot cos^{2}\theta_{W} } \cdot \frac{g^{e}_{i} g^{q}_{j}}{t - M_{Z}^{2}} \; + \; \eta^{eq}_{ij} \label{eq-mt} \end{eqnarray} where $t = -Q^{2}$ is the Mandelstam variable describing the four-momentum transfer between the electron and the quark, $e_{q}$ is the electric charge of the quark in units of the elementary charge and the subscripts $i$ and $j$ label the chiralities of the initial lepton and quark respectively: $i,j=L,R$. $g^{e}_{i}$ and $g^{q}_{j}$ are electroweak couplings of the electron and the quark \begin{eqnarray} g^{f}_{L} & = & I_{3f} - e_{f} \sin^{2}\theta_{W} \label{eq-gdef} \\ g^{f}_{R} & = & \;\;\;\;\; - e_{f} \sin^{2}\theta_{W} \nonumber \end{eqnarray} where $I_{3f}$ is the third component of the SU(2) isospin for the fermion $f$: $f=e,q$. For processes such as $e^{+}e^{-} \rightarrow hadrons$ or $p \bar{p} \rightarrow l^{+} l^{-} X$, a corresponding formula can be written for the $e^{+}e^{-} \rightarrow q \bar{q}$ tree level amplitude: \begin{eqnarray} M^{e_i \bar{e}_j \rightarrow q_i \bar{q}_j}(s) & = & - \frac{4 \pi \alpha_{em} e_{q}}{s} \; + \; \frac{4 \pi \alpha_{em}}{sin^{2}\theta_{W} \cdot cos^{2}\theta_{W} } \cdot \frac{g^{e}_{i} g^{q}_{j}} {s - M_{Z}^{2} + i s \frac{\Gamma_{Z}}{M_{Z}}} \; + \; \eta^{eq}_{ij} \label{eq-ms} \end{eqnarray} where $s$ is the center-of-mass energy squared of the four-fermion reaction. The sign of the contact interaction contribution to the $s$-channel amplitude \eqref{eq-ms} is the same as for the $t$-channel amplitude \eqref{eq-mt}. However, the Standard Model amplitude changes its sign due to the opposite signs of $s$ and $t$ variables. It is therefore important to notice that the resulting sign of the interference terms in the cross-section for $e^{\pm}p$ scattering is different from that in $e^{+}e^{-}$ or $p\bar{p}$ scattering. The contact interaction coupling strength $\eta$ can be related to the mass scale\footnote{Exchanged particle mass or compositeness scale.} ${\cal M}$ of new physics through the formula: \begin{eqnarray} \eta & = & \pm \frac{g_{CI}^{2}}{{\cal M}^{2}} \nonumber \end{eqnarray} where $g_{CI}$ is the unknown coupling strength of new interactions. As the contact interaction contribution always depends on the $g_{CI}$ to ${\cal M}$ ratio, it is convenient to consider the effective mass scale $\Lambda$ defined through the formula: \begin{eqnarray} \eta & = & \pm \frac{4 \pi}{\Lambda^{2}} \nonumber \end{eqnarray} which corresponds to the choice $g^{2}_{CI}=4\pi$. \subsection{General Model} In the most general case, vector contact interactions are described by 4 independent couplings for every lepton-quark pair. With only 2 lepton ($e$ and $\mu$) and 5 quark flavours (i.e. neglecting $t$ quark contribution), we still have 40 independent couplings. It would be very difficult, if not impossible, to consider the model with 40 free parameters. However, some of these parameters (couplings) are weakly constrained by existing experimental data. To reduce the number of the free model parameters, weakly constrained couplings can be either neglected or additionally constrained by relating them to some other couplings. Most of existing experimental data is sensitive predominantly to electron-up and electron-down quark couplings. Therefore, the first model considered in this analysis is the one assuming that these 8 couplings ($\eta^{ed}_{LL}$, $\eta^{ed}_{LR}$, $\eta^{ed}_{RL}$, $\eta^{ed}_{RR}$, $\eta^{eu}_{LL}$, $\eta^{eu}_{LR}$, $\eta^{eu}_{RL}$, $\eta^{eu}_{RR}$) can vary independently, whereas other couplings (for $s$, $c$, $b$, $t$ quarks and/or $\mu$, $\tau$ leptons) are assumed to vanish. This case will be referred to as {\bf the general model}. The other possibility is to impose additional relations between couplings. The common choice is to assume lepton universality: \begin{eqnarray} \eta^{eq}_{ij} & = \eta^{\mu q}_{ij} & = \eta^{\tau q}_{ij} \label{eq-lu} \end{eqnarray} and quark family universality: \begin{eqnarray} \eta^{eu}_{ij} & = \eta^{ec}_{ij} & = \eta^{et}_{ij} \label{eq-qu} \\ \eta^{ed}_{ij} & = \eta^{es}_{ij} & = \eta^{eb}_{ij} \nonumber \end{eqnarray} Lepton universality allows us to include data on muon pair production at the Tevatron (see section \ref{sec-dy}), whereas assuming quark family universality significantly improves the constraints which we can obtain from LEP2 measurements (see section \ref{sec-lep}). As a result, experimental constraints on contact interactions can be significantly improved without increasing the number of free model parameters. The model assuming relations \eqref{eq-lu} and \eqref{eq-qu} will be referred to as {\bf the model with family universality}. \subsection{$SU(2)_{L} \times U(1)_{Y}$ Universality} \label{sec-su2} Another commonly used assumption about lepton-quark contact interactions is that they satisfy the $SU(2)_{L} \times U(1)_{Y}$ gauge invariance of the Standard Model. Assuming that left-handed electrons and quarks belong to $SU(2)_{L}$ doublets and that the contact interaction Lagrangian \eqref{eq-lagr} respects the $SU(2)_{L}$ symmetry implies a relation between contact terms involving left-handed $u$ and $d$ quarks \cite{ciglob}: \begin{eqnarray} \eta^{eu}_{RL} & = & \eta^{ed}_{RL} \nonumber \end{eqnarray} which reduces the number of free model parameters from 8 to 7. The $SU(2)_{L} \times U(1)_{Y}$ also relates $eeqq$ contact interaction couplings with those of $\nu \nu qq$ interactions \begin{eqnarray} \eta^{\nu u}_{LL} & = & \eta^{ed}_{LL} \label{eq-etanu} \\ \eta^{\nu d}_{LL} & = & \eta^{eu}_{LL} \nonumber \\ \eta^{\nu u}_{LR} & = & \eta^{eu}_{LR} \nonumber \\ \eta^{\nu d}_{LR} & = & \eta^{ed}_{LR} \nonumber \end{eqnarray} This allows us to use, in the study of $eeqq$ contact interactions, additional data on NC neutrino scattering (see section \ref{sec-le}). Moreover, assuming the $SU(2)_{L} \times U(1)_{Y}$ universality introduces a related contact interaction term in the Charged Current process $eq \rightarrow \nu q'$. The coupling constant for the induced Charged Current contact interaction is \begin{eqnarray} \eta^{CC} \equiv \eta^{eu \nu d} & = & \eta^{ed}_{LL} - \eta^{eu}_{LL} \label{eq-cc} \end{eqnarray} This relation allows us to use, in the study of Neutral Current contact interaction, also data from Charged Current processes (see section \ref{sec-hera} and \ref{sec-le}). The model assuming the $SU(2)_{L} \times U(1)_{Y}$ universality will be referred to as {\bf the SU(2) model}. In order to reduce the number of models, the SU(2) models considered in this analysis always assume lepton and quark family universality. \subsection{One-parameter models} Using data from a single experiment it is mostly not possible to put significant constraints on contact interaction scales in the general case. Therefore it is a common practise to consider particular models, which assume fixed relations between the separate couplings, reducing the number of free parameters to one. For example, the so called vector-vector model assumes that all couplings are equal: \begin{eqnarray} \eta^{ed}_{LL} = \eta^{ed}_{LR} = \eta^{ed}_{RL} = \eta^{ed}_{RR} = \eta^{eu}_{LL} = \eta^{eu}_{LR} = \eta^{eu}_{RL} = \eta^{eu}_{RR} & \equiv & \eta_{VV} \nonumber \end{eqnarray} Mass scale limits obtained in one-parameter models are, artificially, much stronger than in the general model. They will be considered in this analysis to allow comparison with other results. The relations between couplings assumed for different models are listed in Table \ref{tab-models}\cite{zeusci}. It should be noticed that all one-parameter models considered assume \begin{eqnarray} \eta^{eq}_{LL} + \eta^{eq}_{LR} - \eta^{eq}_{RL} - \eta^{eq}_{RR} & = & 0 \nonumber \end{eqnarray} for $q=u,d$, to avoid strong limits coming from atomic parity violation measurements (see section \ref{sec-le}). For all one-parameter models quark and lepton family universality is assumed. The results obtained both with and without imposing the $SU(2)_{L} \times U(1)_{Y}$ universality are presented, except for the U2, U4 and U6 models, which violate it explicitly ($\eta^{eu}_{RL} \ne \eta^{ed}_{RL}$). \begin{table}[tp] \begin{center} \begin{tabular}{crrrrrrrr} \hline\hline\hline\noalign{\smallskip} Model & $\eta^{ed}_{LL}$ & $\eta^{ed}_{LR}$ & $\eta^{ed}_{RL}$ & $\eta^{ed}_{RR}$ & $\eta^{eu}_{LL}$ & $\eta^{eu}_{LR}$ & $\eta^{eu}_{RL}$ & $\eta^{eu}_{RR}$ \\ \hline\hline\hline\noalign{\smallskip} VV & $+\eta$& $+\eta$& $+\eta$& $+\eta$& $+\eta$& $+\eta$& $+\eta$& $+\eta$\\ AA & $+\eta$& $-\eta$& $-\eta$& $+\eta$& $+\eta$& $-\eta$& $-\eta$& $+\eta$\\ VA & $+\eta$& $-\eta$& $+\eta$& $-\eta$& $+\eta$& $-\eta$& $+\eta$& $-\eta$\\ \hline\noalign{\smallskip} X1 & $+\eta$ & $-\eta$ & & & $+\eta$ & $-\eta$ & & \\ X2 & $+\eta$ & & $+\eta$ & & $+\eta$ & & $+\eta$ & \\ X3 & $+\eta$ & & & $+\eta$ & $+\eta$ & & & $+\eta$ \\ X4 & & $+\eta$ & $+\eta$ & & & $+\eta$ & $+\eta$ & \\ X5 & & $+\eta$ & & $+\eta$ & & $+\eta$ & & $+\eta$ \\ X6 & & & $+\eta$ & $-\eta$ & & & $+\eta$ & $-\eta$ \\ \hline\noalign{\smallskip} U1 & & & & & $+\eta$ & $-\eta$ & & \\ U2 & & & & & $+\eta$ & & $+\eta$ & \\ U3 & & & & & $+\eta$ & & & $+\eta$ \\ U4 & & & & & & $+\eta$ & $+\eta$ & \\ U5 & & & & & & $+\eta$ & & $+\eta$ \\ U6 & & & & & & & $+\eta$ & $-\eta$ \\ \hline\hline\hline\noalign{\smallskip} \end{tabular} \end{center} \caption{Relations between couplings for the one-parameter models considered in this paper.} \label{tab-models} \end{table} \section{Experimental Data} \label{sec-data} In this section the data used to constrain contact interaction model are presented. For each measurement, the formula describing the possible influence of the new couplings on the measured quantities is given. Description of the statistical methods used to interpret the data will be presented in the section \ref{sec-method}. \subsection{High-$Q^{2}$ DIS at HERA} \label{sec-hera} Used in this analysis are the latest data on high-$Q^{2}$ $e^{+}p$ NC DIS from both H1 \cite{h1_98} and ZEUS \cite{zeus98}, corresponding to integrated data luminosities of 37 and 47$pb^{-1}$, respectively. Older results from $e^{-}p$ NC DIS scattering \cite{h1ep,zeusep} are also used, although the influence of these data is marginal. For models with the $SU(2)_{L} \times U(1)_{Y}$ universality, as mentioned in section \ref{sec-model}, data on $e^{+}p$ CC DIS \cite{h1_98,zeuscc} are also included in the fit. HERA experiments quote their high-$Q^{2}$ DIS results in terms of numbers of events and/or cross-sections\footnote{If not given, the number of events can be estimated from the cross-section value assuming that the statistical error quoted corresponds to the Poisson error on the number of measured events, $\sigma_{N} = \sqrt{N}$.} measured in bins of $Q^{2}$. For simplicity let us consider a single $Q^{2}$ bin ranging from $Q^{2}_{min}$ to $Q^{2}_{max}$. Assume that $n_{SM}$ events are expected from the Standard Model. The leading order doubly-differential cross-section for positron-proton NC DIS ($e^{+}p \rightarrow e^{+} X$) can be written as \cite{ciglob} \begin{eqnarray} \frac{d^{2}\sigma^{LO}}{dx dQ^{2}} & = & \frac{1}{16\pi} \sum_{q} q(x) \left\{ |M^{eq}_{LR}|^{2} + |M^{eq}_{RL}|^{2} + (1-y)^{2} \left[ |M^{eq}_{LL}|^{2} + |M^{eq}_{RR}|^{2} \right] \right\} \; + \nonumber \\ & & ~~~~~~~~~~\bar{q}(x)\left\{ |M^{eq}_{LL}|^{2} + |M^{eq}_{RR}|^{2} + (1-y)^{2} \left[ |M^{eq}_{LR}|^{2} + |M^{eq}_{RL}|^{2} \right] \right\} \nonumber \end{eqnarray} where $x$ is the Bjorken variable, describing the fraction of proton momentum carried by a quark (antiquark), $q(x)$ and $\bar{q}(x)$ are the quark and antiquark momentum distribution functions in the proton and $M^{eq}_{ij}$ are the scattering amplitudes of equation \eqref{eq-mt}, which can include contributions from contact interactions described by a set of couplings $\vec{\eta}$. The cross-section (including the contribution from contact interactions), integrated over the $x$ and $Q^{2}$ range of an experimental $Q^{2}$ bin is \begin{eqnarray} \sigma^{LO} ( \vec{\eta} ) & = & \bigint{Q^{2}_{min}}{Q^{2}_{max}} dQ^{2} \bigint{\frac{Q^{2}}{s \cdot y_{max}}}{1} dx \; \frac{d^{2}\sigma^{LO} ( \vec{\eta}) }{dx dQ^{2}} \label{eq-intdis} \end{eqnarray} where $y_{max}$ is the upper limit on reconstructed Bjorken variable $y$ imposed in the analysis\footnote{For the NC DIS analysis H1 uses $y_{max}=0.9$, whereas ZEUS uses $y_{max}=0.95$. For CC DIS analysis both experiments use $y_{max}=0.9$.}. The number of events expected from the Standard Model with contact interaction contributions can now be calculated as: \begin{eqnarray} n(\vec{\eta}) & = & n_{SM} \cdot \left( \frac{\sigma^{LO}(\vec{\eta})}{\sigma^{LO}_{SM}} \right) \label{eq-neta} \end{eqnarray} where $\sigma^{LO}_{SM}$ is the Standard Model cross-section calculated with formula \eqref{eq-intdis} (setting $\vec{\eta}=\vec{0}$). Leading-order expectations of the contact interaction models are used to rescale the Standard Model prediction $n_{SM}$ coming from detailed experiment simulation. This accounts not only for different experimental effects, but also for higher order QCD and electroweak corrections. Validity of this approach is discussed in section \ref{sec-method}. \subsection{Drell-Yan lepton pair production at the Tevatron} \label{sec-dy} Used in this analysis are data on Drell-Yan lepton pair production from the CDF \cite{dy_cdf} and D0 \cite{dy_d0} experiments. Both experiments present numbers of measured high-mass electron pairs ($p \bar{p} \rightarrow e^{+} e^{-}\;X$). CDF also presents results on muon pair production ($p \bar{p} \rightarrow \mu^{+} \mu^{-}\;X$), which are used in the case of models with family universality (see section \ref{sec-model}). The leading order cross-section for lepton pair production in $p \bar{p}$ collisions is \begin{eqnarray} \frac{d^{2}\sigma^{LO}}{dM_{ll} dY} & = & \frac{M_{ll}^{3}}{72\pi s} \sum_{q} q(x_{1}) q(x_{2}) \sum_{i,j=L,R} |M^{eq}_{i j}|^{2} \nonumber \end{eqnarray} where $M_{ll}$ is the invariant mass of lepton pair, $Y$ is the rapidity of the lepton pair center-of-mass frame, $x_{1}$ and $x_{2}$ are the fractions of proton and antiproton momenta carried by the annihilating quarks. The scattering amplitudes $M^{eq}_{i j}$ and the parton density functions are calculated for scale \begin{eqnarray} \hat{s} & = & x_{1} x_{2} s \nonumber \end{eqnarray} where $s$ is the total proton-antiproton center of mass energy squared. The cross-section corresponding to the $M_{ll}$ range from $M_{min}$ to $M_{max}$ is calculated as \begin{eqnarray} \sigma^{LO} ( \vec{\eta} ) & = & \bigint{M_{min}}{M_{max}} dM_{ll} \bigint{-Y_{max}}{Y_{max}} dY \; A_{ll}(Y) \cdot \frac{d^{2}\sigma ( \vec{\eta}) }{dM_{ll} dY} \label{eq-intdy} \end{eqnarray} where $Y_{max}$ is the upper limit on the rapidity of the produced lepton pair: \begin{eqnarray} Y_{max} & = & \ln \frac{\sqrt{s}}{M_{ll}} \; , \nonumber \end{eqnarray} and $A_{ll}(Y)$ is the acceptance function, resulting from the integration over the lepton pair production angle in the center of mass system, with angular detector coverage taken into account. The cross-section calculated with equation \eqref{eq-intdy} is used to calculate the number of events expected from the Standard Model with contact interaction contributions using formula \eqref{eq-neta}. \subsection{Measurements from LEP} \label{sec-lep} Many measurements at LEP are sensitive to different kinds of "new physics". The $eeqq$ contact interactions can be directly tested in the measurement of the total cross-section for $e^{+} e^{-} \rightarrow q \bar{q}$. Using flavour tagging techniques, additional constraints can be obtained from the measurement of the heavy quark decay fractions $R_{b}$ and $R_{c}$, and of the forward-backward asymmetries $A^{q}_{FB}$ of $q\bar{q}$ events. The leading order formula for the total quark pair production cross-section $e^{+} e^{-} \rightarrow q \bar{q}$, at the total electron-positron center of mass energy squared $s$, is \begin{eqnarray} \sigma^{LO}(s) & = & \frac{s}{16\pi} \sum_{q} \sum_{i,j=L,R} |M^{eq}_{i j}|^{2} \label{eq-lep} \end{eqnarray} where $M^{eq}_{ij}$ are the scattering amplitudes described by equation \eqref{eq-ms}, including contributions from contact interaction couplings $\vec{\eta}$. For comparison with measured experimental values, the leading order contact interaction cross-sections are rescaled using the expected Standard Model cross-section $\sigma^{SM}(s)$ quoted by experiments: \begin{eqnarray} \sigma(s,\vec{\eta}) & = & \sigma^{SM}(s) \cdot \left( \frac{\sigma^{LO}(s,\vec{\eta})}{\sigma^{LO}(s,0)} \right) \label{eq-seta} \end{eqnarray} where $\sigma^{LO}(s,0)$ is the leading-order Standard Model cross-section ($\vec{\eta}=\vec{0}$), calculated with equation \eqref{eq-lep}. This takes into account possible experimental effects and higher order QCD and electroweak corrections (for discussion see section \ref{sec-method}). All four LEP experiments have recently presented data on $\sigma_{had}$ for center-of-mass energies up to 189~GeV \cite{aleph,delphi,delphi2,l3,opal}. The sensitivity of the total hadronic cross-section to the contact interaction coupling strength $\vec{\eta}$ is limited by the fact that the interference terms in the quark-pair production cross-sections have opposite signs for up-type and down-type quarks. In the total cross-section, summed over all quark flavours\footnote{Production of the $t$ quark is taken into account only for $\sqrt{s} >$ 350 GeV.}, these terms tend to compensate each other. However, if this is the case, the fraction of events produced with the given quark-pair flavour turns out to be very sensitive to the contact interaction couplings. Using different flavour tagging techniques, cross-sec\-tions for $b\bar{b}$ and $c\bar{c}$ pair production and the corresponding fractions $R_{b}$ and $R_{c}$ can be measured. Although the limited tagging efficiency and purity significantly affects the measurement, useful constraints on contact interaction couplings can be extracted. Used in this analysis are results on $R_{b}$ coming from {\sc Aleph}\cite{aleph,alephrb}, {\sc Delphi} \cite{delphi} and {\sc Opal} \cite{opalrb} as well as {\sc Delphi} results on $R_{c}$ \cite{delphi}. The contact interaction contribution to the scattering amplitude affects also the observed forward-backward asymmetry of $q\bar{q}$ events. In the leading order the forward-backward asymmetry can be calculated as \begin{eqnarray} A^{q}_{FB}(s) & = & \frac{3}{4} \cdot \frac{ |M^{eq}_{LL}|^{2} - |M^{eq}_{LR}|^{2} - |M^{eq}_{RL}|^{2} + |M^{eq}_{RR}|^{2} }{ |M^{eq}_{LL}|^{2} + |M^{eq}_{LR}|^{2} + |M^{eq}_{RL}|^{2} + |M^{eq}_{RR}|^{2} } \nonumber \end{eqnarray} where the factor $\frac{3}{4}$ corresponds to the integration over the full angular range\footnote{For results which are based on the sample of events selected with $| \cos \theta | < 0.9$, this factor is reduced to 0.70866... }. Constraints upon the forward-backward asymmetries $A^{q}_{FB}$ are obtained using a jet charge technique. After clustering all the events into two jets, the jet charge $Q_{jet}$ of each jet can be determined from the momentum weighted sum over all charged tracks in the jet. The sign of $Q_{jet}$ coincides with the charge of produced quark in about 70\% of events. The forward-backward asymmetry for the selected sample of events (e.g. $b$-tagged events) can be extracted in two ways. The method used by {\sc Aleph} is based on the measurement of the mean charge difference between the forward and backward jets $\langle Q_{FB} \rangle = \langle Q^{F}_{jet} \rangle - \langle Q^{B}_{jet} \rangle $. {\sc Delphi} and {\sc Opal} extract $A^{q}_{FB}$ from the angular distribution of jets with well defined sign. In both cases, the measured asymmetry depends on the parton-level asymmetries $A^{q}_{FB}$ and on the quark content of the selected sample. As the up-type and down-type quarks have charges of opposite signs, the measured asymmetry is very sensitive to the relative contribution of different quark flavours. Even if we measure the asymmetry for the flavour-tagged sample, the selected sample of events is always contaminated by other quark flavours (e.g. a $b$-tagged sample always contains a fraction of $c\bar{c}$ events) and the measured value depends strongly on the quark production fractions (e.g. $R_{b}$ and $R_{c}$). This is the reason why the measurement of the forward-backward asymmetry is very sensitive to the contact interaction couplings. Used in this analysis are the measurements of forward-backward asymmetry for the $b$-tagged events \cite{aleph,delphi,opalrb}, $c$-tagged events \cite{delphi} and anti-tagged events \cite{aleph,delphi}. \subsection{Data from low energy experiments} \label{sec-le} The low energy data are included in the present analysis in the manner which follows closely the approach presented in \cite{ciglob,cile}. Therefore only basic assumptions are listed here and technical details are omitted. In case of the general contact interaction model the following constraints from low energy experiments are considered: \begin{itemize} \item Atomic Parity Violation (APV) \\ The Standard Model predicts parity non-conservation in atoms caused (in lowest order) by the $Z^{\circ}$ exchange between electrons and quarks in the nucleus. Experimental results on parity violation in atoms are given in terms of the weak charge $Q_{W}$ of the nuclei. A very precise determination of $Q_{W}$ for Cesium atoms was recently reported \cite{apvcs}. The experimental result differs from the Standard Model prediction \cite{apvth,rpp} by: \begin{eqnarray} \Delta Q_{W}^{Cs} \equiv Q_{W}^{meas} - Q_{W}^{SM} & = & 0.71 \pm 0.84 \nonumber \end{eqnarray} Corresponding results have also been obtained for thallium \cite{apvtl,rpp}: \begin{eqnarray} \Delta Q_{W}^{Tl} & = & 1.9 \pm 3.6 \; . \nonumber \end{eqnarray} These measurements are used to place limits on contact interaction contributions to $Q_{W}$: \begin{eqnarray} \Delta Q_{W}(\vec{\eta}) & = & \frac{2Z+N}{\sqrt{2} G_{F}} \left(\eta^{eu}_{LL}+\eta^{eu}_{LR}-\eta^{eu}_{RL}-\eta^{eu}_{RR} \right) \nonumber \\ & + & \frac{Z+2N}{\sqrt{2} G_{F}} \left(\eta^{ed}_{LL}+\eta^{ed}_{LR}-\eta^{ed}_{RL}-\eta^{ed}_{RR} \right) \nonumber \end{eqnarray} \item electron-nucleus scattering \\ The limits on possible contact interaction contributions to electron-nucleus scattering at low energies can be extracted from the polarisation asymmetry measurement \begin{eqnarray} A & = & \frac{d\sigma_{R}-d\sigma_{L}}{d\sigma_{R}+d\sigma_{L}} \nonumber \end{eqnarray} where $d\sigma_{L(R)}$ denotes the differential cross-section of left- (right-) handed electron scattering. Polarisation asymmetry directly measures the parity violation resulting from the interference between the weak ($Z^{\circ}$ exchange) and the electro-magnetic ($\gamma$ exchange) scattering amplitudes. For isoscalar targets, taking into account valence quark contributions only, the polarisation asymmetry for elastic electron scattering is \begin{eqnarray} A_{el} & = & -\frac{3 \sqrt{2} G_{F} Q^{2}}{20 \pi \alpha_{em} } \left[ 2 \left( g^{u}_{L} + g^{u}_{R} \right) - \left( g^{d}_{L} + g^{d}_{R} \right) \right] \nonumber \end{eqnarray} where $Q^{2}$ is the four-momentum transfer and $g^{q}_{i}$ are quark electroweak couplings, as introduced in equation \eqref{eq-gdef}. Contact interactions modify the effective quark electroweak coupling \begin{eqnarray} \bigmod{g^{q}_{i}}{ef\!f} & = & g^{q}_{i} - \frac{\eta^{e q}_{Li}}{2\sqrt{2} G_{F}} \label{eq-geff} \end{eqnarray} The constraints used in this analysis come from the SLAC $e$D experiment \cite{slac}, the Bates $e$C experiment \cite{bates} and the Mainz experiment on $e$Be scattering \cite{mainz}. In case of models with family universality also data from the $\mu^{\pm}$C experiment at CERN\cite{cern} are included\footnote{The constraints from the $\mu^{\pm}$C experiment result from the comparison of $\mu^{+}_{L} N$ and $\mu^{-}_{R} N$ cross sections.}. \end{itemize} In case of the SU(2) models additional constraints come from: \begin{itemize} \item neutrino-nucleus scattering \\ Constraints on the couplings of quarks to the $Z^{\circ}$ and/ or additional $\nu \nu qq$ contact interactions (related to $eeqq$ CI, as described in section \ref{sec-su2}) can also be derived from the precise measurement of the ratio of Neutral Current to Charged Current neutrino-nucleon scattering cross sections \begin{eqnarray} R^{\nu} & = & \frac{\sigma^{\nu N}_{_{NC}}}{\sigma^{\nu N}_{_{CC}}} \; . \nonumber \end{eqnarray} However, when using constraints on $g^{q}_{i}$ resulting from measurement of $R^{\nu}$, one also has to take into account that possible contact interaction contribution affects not only the Neutral Current but also the Charged Current scattering cross-section (see section \ref{sec-su2}). Therefore the quark electroweak coupling extracted from $R^{\nu}$ measurements should be expressed as\footnote{This correction seems to be missing in \cite{ciglob,cile}.} \begin{eqnarray} \bigmod{g^{q}_{i}}{meas} & = & \frac{g^{q}_{i} - \frac{\eta^{\nu q}_{Li}}{2\sqrt{2} G_{F}}} {1 - \frac{\eta^{CC}}{4\sqrt{2} G_{F}}} \nonumber \end{eqnarray} It is important to notice that $\eta^{e q}_{Li}$ entering formula \eqref{eq-geff} has been replaced here by $\eta^{\nu q}_{Li}$. This is because the effective $g^{u}_{L}$ and $g^{d}_{L}$ couplings measured in neutrino scattering are sensitive to ``flavour crossed'' contact interaction couplings $\eta^{ed}_{LL}$ and $\eta^{eu}_{LL}$ respectively, which results from relations \eqref{eq-etanu}. Experimental constraints on $R^{\nu}$ come mainly from muon-neutrino experiments. Assuming lepton and quark family universality, the following measurements of $g^{q}_{i}$ from $\nu_{\mu} N$ scattering are used: the results compiled by Fogi and Haidt \cite{fh} and the recent constraints from CCFR\cite{ccfr} and NuTeV\cite{nutev}. \item lepton-hadron universality of weak Charged Currents \\ Charged Current contact interactions which are induced by $SU(2)_{L} \times U(1)_{Y}$ universality (see equation \eqref{eq-cc}) would also affect the measurement of $V_{ud}$ element of the Cabibbo-Kobayashi-Maskawa (CKM) matrix, leading to the effective violation of unitarity \cite{cickm,cickm2}. The current experimental constraint is \cite{rpp} \begin{eqnarray} |V_{ud}|^{2} + |V_{us}|^{2} + |V_{ub}|^{2} & = & 0.9969 \pm 0.0022 \nonumber \end{eqnarray} whereas the expected contribution from the contact interaction is \begin{eqnarray} \bigmod{V_{ud}}{meas} & = & V_{ud}^{SM} \cdot \left( 1 - \frac{\eta^{CC}}{4\sqrt{2}G_{F}} \right) \nonumber \end{eqnarray} \item electron-muon universality \\ In the similar way Charged Current contact interactions would also lead to effective violation of $e$-$\mu$ universality in charged pion decay \cite{cickm}. The current experimental value of $R=\Gamma(\pi^{-} \rightarrow e \bar{\nu})/ \Gamma(\pi^{-} \rightarrow \mu \bar{\nu})$ is \cite{emu} \begin{eqnarray} \frac{R_{meas}}{R_{SM}} & = & 0.9966 \pm 0.030 \nonumber \end{eqnarray} whereas the expected contribution from the contact interaction is \begin{eqnarray} \bigmod{R}{meas} & = & R_{SM} \cdot \left( 1 - \frac{\eta^{CC}}{4\sqrt{2}G_{F}} \right)^{2} \nonumber \end{eqnarray} \end{itemize} It is interesting to notice, that data in Charged Current sector may point to a slight violations of the unitarity of the CKM matrix and of the $e$-$\mu$ universality. Both measurements are consistent with the presence of CC contact interactions with a mass scale of the order of 10~TeV. The combined significance of these two results is about 1.8$\sigma$, but it has a considerable influence on global analysis results for the SU(2) model. \section{Analysis method} \label{sec-method} \rightmark The aim of this study is to find the allowed ranges for contact interaction couplings within the different models considered. To do so, the probability function in the coupling space, \begin{eqnarray} {\cal P}(\vec{\eta}) & \sim & \prod_{i} P_{i}(\vec{\eta}), \label{eq-prob} \end{eqnarray} is calculated. In \eqref{eq-prob}, the product runs over all experimental data $i$ and $\vec{\eta}$ represents the set of free parameters for a given model (one or many). This section describes how the probability function is defined and which corrections are included to take into account experimental conditions. \subsection{Statistical errors} \label{sec-met-stat} All experimental data used in this analysis can be divided into two classes. \begin{enumerate} \item For experiments in which a result can be presented as a single number with an error which is considered to reflect a Gaussian probability distribution, the constraints on the contact interaction couplings can be usually expressed using the equation \begin{eqnarray} F(\vec{\eta}) & = & \Delta A \; \pm \; \sigma_{A} \nonumber \end{eqnarray} where $\Delta A$ is the difference between the measured value and the Standard Model prediction, and $F(\vec{\eta})$ is the expected contact interaction contribution to the measured value of A. The resulting probability function can be written as \begin{eqnarray} P_{i}(\vec{\eta}) & \sim & exp \left( -\frac{1}{2} \frac{(F(\vec{\eta}) - \Delta A)^{2}}{\sigma_{A}^{2}} \right) \label{eq-gauss} \end{eqnarray} reflecting the definition of the Gaussian error $\sigma_{A}$. This approach is used for all low energy data as well as for the LEP hadronic cross-section measurements. \item On the other hand, when the experimentally measured quantity is the number of events of a particular kind (e.g. HERA high-$Q^{2}$ events or Drell-Yan lepton pairs at the Tevatron), and especially when this number is small, the probability is better described by the Poisson distribution \begin{eqnarray} P_{i}(\vec{\eta}) & \sim & \frac{ n(\vec{\eta})^{N} \cdot exp( -n(\vec{\eta})) }{ N ! } \label{eq-poisson} \end{eqnarray} where $N$ and $n(\vec{\eta})$ are the measured and expected number of events in a given experiment, respectively, and $n(\vec{\eta})$ takes into account a possible contact interaction contribution. This approach has been used for HERA and the Tevatron data. \end{enumerate} \subsection{Systematic errors} For low energy data the total measurement error can be used in formula \eqref{eq-gauss} taking into account both statistical and systematic errors. For collider data, formula \eqref{eq-gauss} or \eqref{eq-poisson} is used to take into account the statistical error of the measurement only. As for the systematic errors, it is assumed that within a given data set (e.g. $e^{+}p$ NC DIS data from ZEUS ) they are correlated to 100\%. This seems to be a much better approximation of the experimental conditions than assuming that systematic errors are uncorrelated\footnote{Unfortunately the experiments do not publish the correlation matrix for their systematic errors so these are the only possible choices.}. In fact most of the contributing systematic uncertainties at HERA are highly correlated between different $Q^{2}$ bins, as for example energy scale uncertainty or the luminosity measurement. For each data set, a common systematic shift parameter $\delta$ has been introduced to describe the possible variation of event numbers expected at HERA or the Tevatron, or cross-sections predicted at LEP, due to systematic error: \begin{eqnarray} n_{SM} & = & \bar{n}_{SM} \; + \; \delta \cdot \sigma_{n}^{sys} \nonumber \\ {\rm or }\;\;\;\;\;\;\;\;\; \sigma^{SM} & = & \bar{\sigma}_{SM} \; + \; \delta \cdot \sigma_{\sigma}^{sys} \nonumber \end{eqnarray} where $\bar{n}_{SM}$ ($\bar{\sigma}_{SM}$) is the nominal expectation from the Standard Model and $\sigma_{n}^{sys}$ ($\sigma_{\sigma}^{sys}$) is the total systematic uncertainty attributed to this number. Parameters $\delta$ can been treated as additional free parameters when maximising the overall model probability ${\cal P}(\vec{\eta})$. When doing so, normal probability distributions for parameters $\delta$ are included in the definition \eqref{eq-prob} of the probability function\footnote{This corresponds to the assumption that systematic errors are described by the Gaussian probability distribution}. \subsection{Migration corrections} Equation \eqref{eq-neta} introduced in section \ref{sec-data} takes into account experimental conditions at HERA and the Tevatron. The number of events expected with contact interaction contribution is calculated by rescaling the Standard Model prediction $n_{SM}$ coming from the detailed experiment simulation. However, this is only an approximation based on the assumption that the acceptance for contact interaction events is the same as for standard NC DIS events. Although the detection efficiency for given ($x$,$Q^{2}$) or ($M_{ll}$,$Y$) is always the same (as we have the same final state), the distribution of events in the kinematic plane in the presence of the contact interactions can differ significantly. This can affect the measurement due to the finite $Q^{2}$ or $M_{ll}$ resolution. To take this effect into account a dedicated migration correction is introduced. The DIS cross-section in the $Q^{2}$ bin from $Q^{2}_{min}$ to $Q^{2}_{max}$ is calculated by the following extension of eq. \eqref{eq-intdis}: \begin{eqnarray} \sigma^{LO}_{DIS} ( \vec{\eta} ) & = & \bigint{0}{s} dQ^{2} \cdot S(Q^{2};Q^{2}_{min},Q^{2}_{max},\sigma_{Q^{2}}) \; \bigint{\frac{Q^{2}}{s \cdot y_{max}}}{1} dx \frac{d^{2}\sigma ( \vec{\eta}) }{dx dQ^{2}} \nonumber \end{eqnarray} where $\sigma_{Q^{2}}$ is the $Q^{2}$ resolution, as quoted by experiments, assumed to be constant within the bin. The Drell-Yan cross-section is calculated by the similar extension of eq. \eqref{eq-intdy}: \begin{eqnarray} \sigma^{LO}_{DY} ( \vec{\eta} ) & = & \bigint{0}{\sqrt{s}} dM_{ll} \cdot S(M_{ll};M_{min},M_{max},\sigma_{M}) \; \bigint{-Y_{max}}{Y_{max}} dY \; A_{ll}(Y) \cdot \frac{d^{2}\sigma ( \vec{\eta}) }{dM_{ll} dY} \nonumber \end{eqnarray} where $\sigma_{M}$ is the $M_{ll}$ resolution. The mass resolution has been estimated from the quoted calorimeter energy resolution (for electrons) or tracking momentum resolution (for muon pairs). The acceptance function used in both formula \begin{eqnarray} S(x;a,b,\sigma) & = & \bigint{-\infty}{x} dy \; \frac{1}{\sqrt{2 \pi} \sigma } \left[ exp \left( -\frac{1}{2} \frac{(y-a)^{2}}{\sigma^{2}} \right) \; - \; exp \left( -\frac{1}{2} \frac{(y-b)^{2}}{\sigma^{2}} \right) \right] \nonumber \end{eqnarray} describes the probability that the true value $x$ measured with resolution $\sigma$ will be reconstructed between $a$ and $b$. The migration corrections are important for the muon-pair production results from the Tevatron and for the CC DIS results from HERA. For electron-pair production or for NC DIS results, when the corresponding mass and Q$^{2}$ resolutions are much better, the effects of the migration corrections are very small. The influence of the systematic errors and the introduced $Q^{2}$ smearing on the model probability function ${\cal P}(\vec{\eta})$ has been studied for the ZEUS $e^{+}p$ NC DIS data \cite{zeus98}. The results, in terms of the log-likelihood function, -$\log{\cal P}$, for four chosen one-parameter models, are shown in Figure \ref{fig-llsys}. The applied corrections (mainly the systematic error correction) can have sizable influence on the model probability distribution. Taking into account statistical errors only leads to much narrower probability distribution and gives much stronger constraints. The most prominent effect is observed for the VV model. A narrow probability maximum (minimum of -$\log{\cal P}$ function) observed when only statistical errors are included, becomes wider with a ``shoulder'' on one side when the systematic errors are taken into account. \begin{figure}[tp] \centerline{\resizebox{\figwidth}{!}{% \includegraphics{zarneckifig1.eps} }} \caption{Log-likelihood function -$\log{\cal P}(\eta)$ for ZEUS $e^{+}p$ NC DIS data, for four chosen one-parameter models, as indicated on the plot. The functions are calculated with statistical errors only (dashed line) and with migration and systematic error corrections (solid line).} \label{fig-llsys} \end{figure} The results from this analysis have been compared with the ZEUS results based on full detector simulation \cite{zeusci}. The comparison for the same four one-parameter models is presented in Figure \ref{fig-llcomp}. For some models very good agreement is observed between this analysis and ZEUS results, as can be seen for AA and X1 models. However, for models such as VV or U2, the constraints given by ZEUS are stronger (probability distribution narrower) than the constraints resulting from this analysis. This is due to the fact that the ZEUS analysis takes into account the two-dimensional event distribution in the $(x,y)$ plane, whereas this analysis uses the one-dimensional $Q^{2}$ distribution only. \begin{figure}[tp] \centerline{\resizebox{\figwidth}{!}{% \includegraphics{zarneckifig2.eps} }} \caption{Log-likelihood function -$\log{\cal P}(\eta)$ for ZEUS $e^{+}p$ NC DIS data, for four chosen one-parameter models, as indicated on the plot. The results from this analysis (solid line) are compared with the ZEUS results obtained with full detector simulation (dashed line). } \label{fig-llcomp} \end{figure} \subsection{Radiative corrections} For high-energy data from HERA, LEP and the Tevatron, Standard Model predictions given by experiments are used to rescale leading-order expectations of the contact interaction models (see formula \eqref{eq-neta} and \eqref{eq-seta}). This accounts not only for different experimental effects, but also for higher order QCD and electroweak corrections, including radiative corrections. This approach is reasonable as long as the difference between the corrections for the Standard Model and for the model including contact interactions is negligible. It is natural to assume that this difference should be much smaller than the correction itself. The contribution of radiative corrections to high-Q$^{2}$ DIS at HERA is of the order of 10\%. For high-mass Drell-Yan lepton pair production at the Tevatron it is only about 6\%. Therefore, the possible variation of the radiative corrections for both HERA and Tevatron data have been neglected. The only data where radiative corrections could be significant is the hadronic cross-section measurement at LEP. Most of the events observed at LEP2 are radiative events. This is due to the "radiative escape" to the $Z^{\circ}$ peak. Radiation probability is significantly enhanced as the $e^{+}e^{-}$ annihilation cross-section at $\sqrt{s}=M_{Z}$ is several orders of magnitude higher than at nominal $\sqrt{s}$. The leading order cross-section \eqref{eq-lep} is corrected for radiation effects using the formula \cite{leprad} \begin{eqnarray} \sigma_{rad}(s, \vec{\eta} ) & = & \bigint{s'_{min}}{s} \frac{ds'}{s} \;\;\; G(\frac{s'}{s}) \cdot \sigma^{LO} (s' , \vec{\eta}) \nonumber \end{eqnarray} where integration runs over the center-of-mass energy squared $s'$ of the produced quark pair, and $s'_{min}$ is the minimum value of $s'$ required by the event selection cuts\footnote{Data used in this analysis correspond to $\sqrt{s'/s}>0.9$ (ALEPH) or $\sqrt{s'/s}>0.85$ (DELPHI, L3 and OPAL). This choice significantly reduces possible influence of radiative corrections.}. G(z) is the "radiator function" encapsulating the results of QED virtual and real corrections. Used in this analysis is the approximate formula (based on \cite{leprad,leprad2} ) \begin{eqnarray} G(z) & = & f_{r} \cdot \beta (1-z)^{\beta - 1} \; + \; (1-f_{r}) \cdot \delta (1-z) \nonumber \\[3mm] {\rm where } \;\;\;\;\;\;\;\; \beta & = & 2 \; \frac{\alpha_{em}}{\pi} \; \left( log \frac{s}{m_{e}^{2}} - 1 \right) \nonumber \end{eqnarray} The parameter $f_{r}$ is chosen to reproduce the cross-section ratio for radiative and non-radiative events\footnote{Events with 0.1$<\sqrt{s'/s}<$0.85 and $\sqrt{s'/s}>$0.85 \cite{opal}.}. It turned out that the effect of radiative corrections on the probability function ${\cal P}(\vec{\eta})$ is very small. The resulting limits on contact interaction mass parameters decrease by at most 3\%. \subsection{Probability functions} \label{sec-met-prob} The probability function ${\cal P}(\vec{\eta})$ summarises our current experimental knowledge about possible $eeqq$ contact interactions. It will be used to set limits on contact interaction mass scale parameters and to extract predictions concerning possible future discoveries. It is therefore very important to understand the precise meaning of ${\cal P}(\vec{\eta})$. ${\cal P}(\vec{\eta})$ is {\bf not} a probability distribution of $\vec{\eta}$. A probability distribution should describe the probability of finding a given value of variable. Our situation is different. Function ${\cal P}(\vec{\eta})$ describes the probability that our data come from the model described by the set of couplings $\vec{\eta}$ (see section \ref{sec-met-stat}). It is our data set, which is a variable, and $\vec{\eta}$ is a set of model parameters: they are unknown, but they are fixed. This simple observation has very important implications for this analysis, not only for the limit setting procedure (see next subsection) but also for calculation of model predictions. To set limits on possible deviations from the Standard Model predictions (eg. for NC DIS cross-section at very high-Q$^{2}$ at HERA or for hadronic cross-section at next $e^{+}e^{-}$ collider), we have to consider the probability function $P(r)$, where the cross-section deviation $r$ is defined as \begin{eqnarray} r & = & \frac{\sigma(\vec{\eta})}{\sigma_{SM}} \; = \; R(\vec{\eta}) \nonumber \end{eqnarray} If ${\cal P}(\vec{\eta})$ is taken as probability distribution, then the probability distribution for $r$ should be calculated as \begin{eqnarray} P(r) & = & \int d^{N}\vec{\eta} \;\;\; {\cal P}(\vec{\eta}) \; \delta(r-R(\vec{\eta})) \label{eq-oldprob} \end{eqnarray} where integration is performed over N-dimensional coupling space. This however leads to completely false results, as is demonstrated in appendix \ref{app-toy}. Instead of calculating the probability distribution for $r$ (which is not well defined), we should rather try to find out what is the probability that our data come from the model predicting deviation $r$. This leads to the formula: \begin{eqnarray} P(r) & = & \bigmod{\left<{\cal P}(\vec{\eta}) \right>}{R(\vec{\eta})=r} \nonumber \end{eqnarray} where averaging is necessary, if we want to reduce number of parameters of the probability function (for multi-parameter models). The commonly used assumption in that case, is that $\vec{\eta}$ has flat underlying (prior) distribution\footnote{This corresponds to the assumption, that we would have no preferences for any value of $\vec{\eta}$, if there is no experimental data.}. The formula for $P(r)$ can be then expressed as \begin{eqnarray} P(r) & = & \frac{ \displaystyle \int d^{N}\vec{\eta} \; {\cal P}^{2}(\vec{\eta}) \; \delta(r-R(\vec{\eta})) }{ \displaystyle \int d^{N}\vec{\eta} \; {\cal P}(\vec{\eta}) \; \delta(r-R(\vec{\eta})) } \label{eq-newprob} \end{eqnarray} The formula applies for any variable which can be used as a parameter of the probability function. In this analysis it will also be used to calculate probability functions and to set limits on mass scale parameters corresponding to single couplings in multi-parameter models. As ${\cal P}(\vec{\eta})$ is not the probability distribution it does not satisfy any normalisation condition. Instead it is convenient to rescale the probability function in such a way that its global maximum has the value of 1: \begin{eqnarray} \max_{\vec{\eta}} {\cal P}(\vec{\eta}) & = & 1. \label{eq-pmax} \end{eqnarray} \subsection{Extracting limits} \label{sec-met-lim} After imposing condition \eqref{eq-pmax}, the lower and upper limits on the value of the model parameter $r$ are defined as minimum ($r^{-}$) and maximum ($r^{+}$ ) values satisfying relation \begin{eqnarray} P(r^{-}) & = & 0.05 \nonumber \\ {\rm and} \;\;\;\;\; P(r^{+}) & = & 0.05 \; . \nonumber \end{eqnarray} For any model described by the parameter $r < r^{-}$ or $r > r^{+}$, the probability that our data results from this model is less than 5\% of the maximum probability. This is taken as the definition of the 95\% confidence level (CL) limits. For one-parameter contact interaction models this approach is slightly modified. As models with negative and positive values of $\eta$ are usually considered as independent scenarios (differing by the signs of the interference terms in the cross-section), the upper and lower limits on $\eta$ are calculated using restricted $\eta$ range: \begin{eqnarray} P(\eta^{-}) & = & 0.05 \cdot \max_{\eta < 0} P(\eta) \nonumber \\ {\rm and} \;\;\;\;\; P(\eta^{+}) & = & 0.05 \cdot \max_{\eta > 0} P(\eta) \; . \label{eq-elimit} \end{eqnarray} For one-parameter contact interaction models, or for probability functions related to single couplings in multi-parameter models, the limits on coupling values $\eta^{-}$ and $\eta^{+}$ can be translated into the limits on contact interaction mass scales \begin{eqnarray} \Lambda^{-} & = & \sqrt{\frac{4 \pi}{ - \eta^{-}}} \nonumber \\ \Lambda^{+} & = & \sqrt{\frac{4 \pi}{\eta^{+}}} \nonumber \end{eqnarray} Mass limits commonly used in literature are based on $\eta$ limits defined in a slightly different way. In this paper they will be denoted as $\eta^{--}$ and $\eta^{++}$. Their definition follows from the equations: \begin{eqnarray} \bigint{\eta^{--}}{0} d\eta \; P(\eta) & = & 0.95 \cdot \bigint{-\infty}{0} d\eta \; P(\eta) \nonumber \\ \bigint{0}{\eta^{++}} d\eta \; P(\eta) & = & 0.95 \cdot \bigint{0}{\infty} d\eta \; P(\eta) \label{eq-elimit2} \end{eqnarray} This approach is based on the assumption that $\eta$ has a flat underlying (prior) distribution. In such a case $P(\eta)$ can be treated as the probability distribution for $\eta$. The mass scale limits corresponding to $\eta^{--}$ and $\eta^{++}$ will be denoted as $\Lambda^{--}$ and $\Lambda^{++}$. Although the definition resulting from equation \eqref{eq-elimit} is considered to be more appropriate for this analysis than definition \eqref{eq-elimit2}, the results for both definitions are presented to allow comparison with other results. As definitions \eqref{eq-elimit} and \eqref{eq-elimit2} correspond to the different interpretation of the probability function, they are not expected to give similar results. In fact, the allowed range for parameter $\eta$, calculated with equation \eqref{eq-elimit} is usually about 25\% wider than the one calculated with equation \eqref{eq-elimit2}\footnote{For Gaussian shape of the probability function, $\eta^{-}$ and $\eta^{+}$ correspond to $\pm 2.45 \sigma$ limits, whereas $\eta^{--}$ and $\eta^{++}$ to $\pm 1.96 \sigma$.}. As a result, corresponding mass scale limits $\Lambda^{-}$ and $\Lambda^{+}$ are usually 10 to 15\% smaller than $\Lambda^{--}$ and $\Lambda^{++}$. \section{Results} \label{sec-results} For one-parameter models the analysis has been performed both without and with the additional $SU(2)_{L} \times U(1)_{Y}$ universality assumption (see section \ref{sec-model}). In the latter case data coming from neutrino-nucleus scattering experiments and from different Charged Current processes ( refer section \ref{sec-data}) have been also used to constraint contact interaction couplings. In the following, the models assuming $SU(2)_{L} \times U(1)_{Y}$ symmetry are referred to as SU(2) models. One-parameter models without $SU(2)_{L} \times U(1)_{Y}$ symmetry will be referred to as simple models, to avoid possible confusion. For all one-parameter models quark and lepton family universality is assumed. Using the overall model probability ${\cal P}(\vec{\eta})$, as defined by equation \eqref{eq-prob}, the "best" values of contact interaction couplings (i.e. corresponding to the maximum probability) were found using the MINUIT package \cite{minuit}. The results for one-parameter simple and SU(2) models are presented in Tables \ref{tab-result1f} and \ref{tab-result1t}, respectively. The errors attributed to $\eta$ values correspond to the decrease in the model probability ${\cal P}(\eta)$ by the factor of $\sqrt{e}$. In case of asymmetric errors the arithmetic mean is given. \begin{table}[tp] \begin{center} \begin{tabular}{cr@{~$\pm$}lrrrr} \hline\hline\hline\noalign{\smallskip} Model & \multicolumn{2}{c}{$\eta$} & \multicolumn{4}{c}{Mass scale limits [TeV]} \\ \cline{4-7}\noalign{\smallskip} & \multicolumn{2}{c}{[TeV$^{-2}$]} & $\Lambda^{-}$ & $\Lambda^{+}$ & $\Lambda^{--}$ & $\Lambda^{++}$ \\ \hline\hline\hline\noalign{\smallskip} VV & -0.015 & 0.049 & 9.8 & 10.7 & 11.0 & 11.8 \\ AA & 0.007 & 0.048 & 10.5 & 10.1 & 11.7 & 11.3 \\ VA & 0.049 & 0.143 & 6.6 & 6.2 & 7.3 & 6.9 \\ \hline\noalign{\smallskip} % X1 & 0.014 & 0.073 & 8.7 & 8.1 & 9.6 & 9.2 \\ X2 & -0.011 & 0.075 & 8.2 & 8.4 & 9.2 & 9.4 \\ X3 & -0.003 & 0.051 & 9.9 & 10.2 & 11.1 & 11.4 \\ X4 & -0.113 & 0.138 & 5.7 & 5.2 & 6.4 & 5.3 \\ X5 & -0.079 & 0.132 & 5.9 & 6.4 & 6.6 & 7.0 \\ X6 & -0.013 & 0.147 & 6.2 & 5.8 & 7.0 & 6.4 \\ \hline\noalign{\smallskip} % U1 & -0.059 & 0.104 & 6.4 & 7.7 & 7.3 & 8.4 \\ U2 & -0.065 & 0.082 & 6.9 & 9.1 & 7.8 & 9.9 \\ U3 & -0.044 & 0.053 & 8.5 & 11.7 & 9.6 & 12.7 \\ U4 & -0.136 & 0.166 & 5.1 & 5.5 & 5.7 & 5.8 \\ U5 & -0.093 & 0.092 & 6.4 & 8.8 & 7.2 & 9.5 \\ U6 & 0.115 & 0.128 & 7.0 & 5.6 & 7.4 & 6.3 \\ \hline\hline\hline\noalign{\smallskip} \end{tabular} \end{center} \caption{Coupling values and 95\% CL mass scale limits resulting from fits of one-parameter models {\bf without} $SU(2)_{L} \times U(1)_{Y}$ universality. The errors attributed to $\eta$ values correspond to the decrease in the model probability ${\cal P}(\eta)$ by the factor of $\sqrt{e}$. See text for explanation of the symbols. } \label{tab-result1f} \end{table} For all simple one-parameter models considered couplings are found to be consistent with the Standard Model within $1 \sigma$. The same is true for most SU(2) models. However, for the SU(2) models U1 and U3 the ``best'' coupling values are more than $2 \sigma$ from the Standard Model. These are the only two models which allow for $\eta^{CC} \neq 0$ (i.e. $\eta^{eu}_{LL} \neq \eta^{ed}_{LL} $). The observed deviation from the Standard Model predictions is directly related to $\eta^{CC}$ bounds coming from the unitarity of the CKM matrix and the $e$-$\mu$ universality, as described in section \ref{sec-data}. However, it has to be noticed that other data also do support this effect: the discrepancy observed for the combined data is more significant than for the Charged Current sector only. Although the effect is interesting, the data are still in acceptable agreement with the Standard Model. The probability that our data result from the Standard Model equals 5.7\% and 7.0\% for the U1 and U3 SU(2) models respectively. Assuming that there is no direct evidence for $eeqq$ contact interactions, the limits on single couplings can be calculated. The results on mass scale limits $\Lambda^{-}$, $\Lambda^{+}$, $\Lambda^{--}$ and $\Lambda^{++}$ obtained from fitting one-parameter models are summarised in Tables \ref{tab-result1f} and \ref{tab-result1t}. For simple models mass limits range from 5.1 TeV ($\Lambda^{-}$ for the U4 model) to 11.7 TeV ($\Lambda^{+}$ for the U3 model). Similar limits are obtained for most of the SU(2) models. Only for the U1 and U3 SU(2) models much higher $\Lambda^{+}$ limits of 17.0 and 18.2 TeV are obtained. The probability functions ${\cal P}(\eta)$ for four selected SU(2) models are shown in Figure \ref{fig-prob1}. \begin{table}[tp] \begin{center} \begin{tabular}{cr@{~$\pm$}lrrrr} \hline\hline\hline\noalign{\smallskip} Model & \multicolumn{2}{c}{$\eta$} & \multicolumn{4}{c}{Mass scale limits [TeV]} \\ \cline{4-7}\noalign{\smallskip} & \multicolumn{2}{c}{[TeV$^{-2}$]} & $\Lambda^{-}$ & $\Lambda^{+}$ & $\Lambda^{--}$ & $\Lambda^{++}$ \\ \hline\hline\hline\noalign{\smallskip} % VV & -0.024 & 0.047 & 9.6 & 11.4 & 10.8 & 12.5 \\ AA & -0.010 & 0.047 & 9.9 & 11.1 & 11.1 & 12.3 \\ VA & -0.078 & 0.108 & 6.3 & 8.0 & 7.1 & 8.7 \\ \hline\noalign{\smallskip} % X1 & -0.025 & 0.067 & 8.1 & 9.5 & 9.2 & 10.5 \\ X2 & -0.041 & 0.069 & 7.8 & 9.6 & 8.8 & 10.5 \\ X3 & -0.019 & 0.049 & 9.5 & 11.1 & 10.7 & 12.2 \\ X4 & -0.066 & 0.144 & 6.0 & 5.4 & 6.7 & 5.8 \\ X5 & -0.040 & 0.131 & 6.2 & 6.4 & 7.0 & 7.0 \\ X6 & -0.013 & 0.147 & 6.2 & 5.8 & 7.0 & 6.4 \\ \hline\noalign{\smallskip} % U1 & -0.100 & 0.042 & 7.9 & 17.0 & 8.6 & 17.8 \\ U3 & -0.083 & 0.036 & 8.6 & 18.2 & 9.4 & 19.1 \\ U5 & -0.050 & 0.082 & 7.1 & 8.8 & 8.0 & 9.6 \\ \hline\hline\hline\noalign{\smallskip} \end{tabular} \end{center} \caption{Coupling values and mass scale limits resulting from fits of one-parameter models {\bf with} $SU(2)_{L} \times U(1)_{Y}$ universality. The errors attributed to $\eta$ values correspond to the decrease in the model probability ${\cal P}(\eta)$ by the factor of $\sqrt{e}$. See text for explanation of the symbols. } \label{tab-result1t} \end{table} \begin{figure}[tp] \centerline{\resizebox{\figwidth}{!}{% \includegraphics{zarneckifig3.eps} }} \caption{ Probability functions ${\cal P}(\eta)$ for chosen one-parameter models with $SU(2)_{L} \times U(1)_{Y}$ universality, as indicated on the plot.} \label{fig-prob1} \end{figure} Contribution of different data sets to the mass scale limits presented in Tables \ref{tab-result1f} and \ref{tab-result1t} can be estimated using the probability function. Mass scale limits $\Lambda^{-}$ and $\Lambda^{+}$, derived from coupling limits $\eta^{-}$ and $\eta^{+}$, correspond to the decrease of the global probability ${\cal P}(\eta)$ to 0.05 of its maximum value (see equation \eqref{eq-elimit}). This decrease can be presented as a product of contributions from all data samples. Table \ref{tab-probdata} presents the relative probability changes, calculated separately for different data sets, corresponding to mass scale limits $\Lambda^{-}$ and $\Lambda^{+}$, for different one-parameter SU(2) models. The product of numbers in every row is equal to the factor 0.05 defining the 95\% CL. Numbers close to 1.0 demonstrate that given data set has negligible influence on the considered mass scale limit. The smaller the number, the more sensitive are the data to a given CI model. Numbers greater than 1.0 indicate that the model with mass scale $\Lambda^{-}$ or $\Lambda^{+}$ gives better description of given data set than the ``best'' coupling value resulting from the combined fit. The results presented in Table \ref{tab-probdata} show that in most models the strongest constraints on contact interaction couplings come from LEP data on forward-backward asymmetries $A^{q}_{FB}$ and on quark production ratios $R_{q}$. However, for particular models a significant contribution can result from LEP hadronic cross-section measurements, neutrino-nucleus scattering data, HERA NC DIS data or from data on Charged Current interactions. \renewcommand{\multirowsetup}{\centering} \begin{table}[tp] \begin{tabular*}{\textwidth}{@{}cc@{\extracolsep{4in minus 4in}}cccccccc@{}} \hline\hline\hline\noalign{\smallskip} & Mass & \multicolumn{8}{c}{Relative change in model probability} \\ \cline{3-10}\noalign{\smallskip} Model & scale & HERA & Tevatron & \multicolumn{3}{c}{LEP} & \multicolumn{2}{c}{Low energy NC} & CC \\ \cline{5-7} \cline{8-9}\noalign{\smallskip} & limit & $e^{\pm}p$ NC & Drell-Yan & $\sigma_{had}$ & $R_{q}$ & $A^{q}_{FB}$ & $l^{\pm}N$ & $\nu N$ & data \\ \hline\hline\hline\noalign{\smallskip} \multirow{2}{10mm}{VV} & $\Lambda^{-}$ & 0.697 & 0.887 & 1.344 & 1.665 & 0.028 & 1.000 & 1.300 & \\ & $\Lambda^{+}$ & 1.107 & 0.632 & 0.751 & 0.205 & 0.841 & 1.000 & 0.553 & \\ \noalign{\smallskip} \multirow{2}{10mm}{AA} & $\Lambda^{-}$ & 1.240 & 1.021 & 0.536 & 1.867 & 0.020 & 0.984 & 1.961 & \\ & $\Lambda^{+}$ & 0.774 & 0.835 & 1.451 & 0.208 & 0.722 & 1.013 & 0.350 & \\ \noalign{\smallskip} \multirow{2}{10mm}{VA} & $\Lambda^{-}$ & 1.264 & 0.710 & 1.281 & 1.092 & 0.033 & 0.916 & 1.316 & \\ & $\Lambda^{+}$ & 1.008 & 0.928 & 1.370 & 0.811 & 0.504 & 1.059 & 0.090 & \\ \hline\noalign{\smallskip} \multirow{2}{10mm}{X1} & $\Lambda^{-}$ & 1.214 & 1.013 & 0.694 & 1.746 & 0.017 & 0.962 & 2.060 & \\ & $\Lambda^{+}$ & 0.830 & 0.885 & 1.395 & 0.344 & 0.634 & 1.031 & 0.217 & \\ \noalign{\smallskip} \multirow{2}{10mm}{X2} & $\Lambda^{-}$ & 0.842 & 0.931 & 1.323 & 1.659 & 0.017 & 0.973 & 1.773 & \\ & $\Lambda^{+}$ & 1.079 & 0.748 & 0.878 & 0.327 & 0.660 & 1.021 & 0.320 & \\ \noalign{\smallskip} \multirow{2}{10mm}{X3} & $\Lambda^{-}$ & 1.016 & 1.005 & 0.993 & 1.734 & 0.017 & 0.992 & 1.720 & \\ & $\Lambda^{+}$ & 0.977 & 0.710 & 1.137 & 0.192 & 0.738 & 1.007 & 0.444 & \\ \noalign{\smallskip} \multirow{2}{10mm}{X4} & $\Lambda^{-}$ & 0.331 & 0.762 & 0.318 & 1.021 & 1.414 & 1.018 & 0.424 & \\ & $\Lambda^{+}$ & 0.814 & 0.666 & 0.218 & 0.779 & 0.690 & 0.969 & 0.813 & \\ \noalign{\smallskip} \multirow{2}{10mm}{X5} & $\Lambda^{-}$ & 0.617 & 0.650 & 1.178 & 1.749 & 0.133 & 1.043 & 0.435 & \\ & $\Lambda^{+}$ & 1.142 & 0.516 & 0.884 & 0.064 & 1.411 & 0.949 & 1.114 & \\ \noalign{\smallskip} \multirow{2}{10mm}{X6} & $\Lambda^{-}$ & 0.771 & 0.786 & 1.127 & 0.077 & 0.984 & 0.965 & & \\ & $\Lambda^{+}$ & 1.367 & 0.731 & 0.789 & 1.976 & 0.031 & 1.025 & & \\ \hline\noalign{\smallskip} \multirow{2}{10mm}{U1} & $\Lambda^{-}$ & 1.141 & 0.963 & 0.895 & 1.207 & 0.361 & 0.925 & 0.443 & 0.285 \\ & $\Lambda^{+}$ & 0.933 & 0.954 & 0.753 & 0.903 & 1.174 & 1.028 & 0.323 & 0.212 \\ \noalign{\smallskip} \multirow{2}{10mm}{U3} & $\Lambda^{-}$ & 1.056 & 0.855 & 0.254 & 1.289 & 0.243 & 0.987 & 1.157 & 0.611 \\ & $\Lambda^{+}$ & 0.963 & 0.885 & 0.502 & 0.804 & 1.212 & 1.005 & 0.449 & 0.266 \\ \noalign{\smallskip} \multirow{2}{10mm}{U5} & $\Lambda^{-}$ & 0.724 & 0.849 & 0.447 & 1.535 & 0.561 & 1.084 & 0.195 & \\ & $\Lambda^{+}$ & 1.004 & 0.671 & 0.109 & 0.570 & 1.267 & 0.910 & 1.036 & \\ \hline\hline\hline\noalign{\smallskip} \end{tabular*} \caption{Relative changes in model probabilities calculated for separate data sets (as indicated in the table) corresponding to the decrease in the global probability at mass scale limit ($\Lambda^{-}$ or $\Lambda^{+}$) to 0.05 of its maximum value (for negative or positive couplings respectively). Considered are one-parameter models with family and $SU(2)_{L} \times U(1)_{Y}$ universality. } \label{tab-probdata} \end{table} \begin{figure}[tp] \centerline{\resizebox{\figwidth}{!}{% \includegraphics{zarneckifig4.eps} }} \caption{ Probability functions $P(\eta)$ for single contact interaction couplings (as indicated on the plot) obtained within the general contact interaction model.} \label{fig-prob2} \end{figure} The probability functions for single couplings obtained for the general model are presented in Figure \ref{fig-prob2}. All couplings are consistent with the Standard Model predictions. Results for single couplings obtained for the general model, the model with family universality and the SU(2) model with family universality are summarised in Table \ref{tab-result2}. It has to be stressed that all limits for single couplings are derived without any assumptions concerning the remaining couplings, which corresponds to the definition \eqref{eq-newprob} of a probability function. For this reason most calculated limits are weaker than in case of one-parameter models. The mass limits obtained for the general model range from 2.1 TeV ($\Lambda^{ed\;+}_{RL}$) to 5.1 TeV ($\Lambda^{eu\;-}_{LL}$). For the SU(2) model with family universality, the corresponding numbers are 3.5 and 7.8 TeV for $\Lambda^{ed\;-}_{RR}$ and $\Lambda^{eu\;+}_{LL}$, respectively. \begin{table*}[tp] \begin{tabular*}{\textwidth}{c@{\extracolsep{4in minus 4in}}rlrlrlrlrlrl} \hline\hline\hline\noalign{\smallskip} & \multicolumn{12}{c}{Mass scale limits [TeV] } \\ \cline{2-13}\noalign{\smallskip} & \multicolumn{4}{c}{ } & \multicolumn{4}{c}{Model with} & \multicolumn{4}{c}{SU(2) model with} \\ Coupling & \multicolumn{4}{c}{General model } & \multicolumn{4}{c}{family universality} & \multicolumn{4}{c}{family universality} \\ \cline{2-5} \cline{6-9} \cline{10-13} \noalign{\smallskip} & $\Lambda^{-}$ & $\Lambda^{+}$ & $\Lambda^{--}$ & $\Lambda^{++}$ & $\Lambda^{-}$ & $\Lambda^{+}$ & $\Lambda^{--}$ & $\Lambda^{++}$ & $\Lambda^{-}$ & $\Lambda^{+}$ & $\Lambda^{--}$ & $\Lambda^{++}$ \\ \hline\hline\hline\noalign{\smallskip} $\eta^{ed}_{LL}$ & 3.1 & 3.6 & 3.4 & 4.0 & 4.3 & 5.4 & 4.8 & 6.0 & 7.5 & 6.1 & 8.1 & 6.8 \\ % $\eta^{ed}_{LR}$ & 2.4 & 2.2 & 2.6 & 2.5 & 2.8 & 3.2 & 3.0 & 3.6 & 3.6 & 3.7 & 4.0 & 4.1 \\ % $\eta^{ed}_{RL}$ & 2.6 & 2.1 & 2.8 & 2.4 & 3.9 & 2.8 & 4.4 & 3.1 & 4.5 & 3.9 & 5.0 & 4.4 \\ % $\eta^{ed}_{RR}$ & 2.8 & 2.8 & 3.0 & 3.1 & 3.3 & 3.5 & 3.7 & 3.9 & 3.5 & 5.1 & 3.8 & 5.7 \\ % $\eta^{eu}_{LL}$ & 5.1 & 3.9 & 5.6 & 4.3 & 5.3 & 4.1 & 5.9 & 4.5 & 6.5 & 7.8 & 7.3 & 8.5 \\ % $\eta^{eu}_{LR}$ & 4.1 & 3.3 & 4.3 & 3.7 & 3.7 & 3.5 & 4.1 & 3.9 & 4.9 & 4.7 & 5.4 & 5.3 \\ % $\eta^{eu}_{RL}$ & 3.1 & 3.5 & 3.5 & 3.9 & 3.1 & 3.5 & 3.4 & 3.9 & 4.5 & 3.9 & 5.0 & 4.4 \\ % $\eta^{eu}_{RR}$ & 3.7 & 3.7 & 4.1 & 4.1 & 3.8 & 4.2 & 4.3 & 4.6 & 4.8 & 4.4 & 5.4 & 4.8 \\ \hline\hline\hline\noalign{\smallskip} \end{tabular*} \caption{95\% CL mass scale limits for single couplings, obtained within different models, as indicated in the table. See text for mass scale limits definition.} \label{tab-result2} \end{table*} Single couplings can either increase or decrease cross-section for a given process, as compared with the Standard Model expectations. It is therefore also possible that the influence of the two different couplings compensate each other. Because of that, the limits on mass scales obtained for single couplings does not exclude contact interactions with smaller mass scales. To obtain the most general limit, the eigenvectors of the correlation matrix (obtained from MINUIT from the functional form of ${\cal P}(\vec{\eta})$ in the vicinity of the maximum probability) are considered. In case of the general model the two least constrained linear coupling combinations are \begin{eqnarray} \eta_{1} & = & -0.26 \eta^{ed}_{LL} + 0.84 \eta^{ed}_{LR} +0.15 \eta^{ed}_{RL} + 0.33 \eta^{ed}_{RR} \nonumber \\ & & -0.06 \eta^{eu}_{LL} + 0.11 \eta^{eu}_{LR} -0.10 \eta^{eu}_{RL} + 0.27 \eta^{eu}_{RR} \nonumber \\ {\rm and} \;\;\;\;\;\;\; \eta_{2} & = & +0.20 \eta^{ed}_{LL} + 0.17 \eta^{ed}_{LR} +0.81 \eta^{ed}_{RL} - 0.48 \eta^{ed}_{RR} \nonumber \\ & & +0.10 \eta^{eu}_{LL} - 0.14 \eta^{eu}_{LR} +0.10 \eta^{eu}_{RL} - 0.09 \eta^{eu}_{RR} \nonumber \end{eqnarray} This is in agreement with the observation that the least constrained single couplings are $\eta^{ed}_{LR}$ and $\eta^{ed}_{RL}$, which can also be seen from Table \ref{tab-result2}. The probability functions for $\eta_{1}$ and $\eta_{2}$ are shown in Figure \ref{fig-prob3}. The mass scale limit corresponding to $\eta_{1}$ is\footnote{As the sign of $\eta_{1}$ is arbitrary, only one value is given. It is calculated as min($\Lambda^{+}_{1},\Lambda^{-}_{1}$).} \begin{eqnarray} \Lambda_{1} & = & 2.1 \; \rm TeV \nonumber \end{eqnarray} This limit should be considered to be the most general one, as it is valid for any combination of couplings. This means that any contact interaction with a mass scale below 2.1 TeV is excluded on 95\% CL. On the other hand it also shows that the existing data do not exclude mass scales of the order of 3 TeV. The limits on the mass scale associated with $\eta_{1}$ are summarised in Table \ref{tab-result3}. \begin{table*}[tp] \begin{tabular*}{\textwidth}{c@{\extracolsep{4in minus 4in}}rlrlrlrlrlrl} \hline\hline\hline\noalign{\smallskip} & \multicolumn{12}{c}{Mass scale limits [TeV] } \\ \cline{2-13}\noalign{\smallskip} & \multicolumn{4}{c}{ } & \multicolumn{4}{c}{Model with} & \multicolumn{4}{c}{SU(2) model with} \\ Coupling & \multicolumn{4}{c}{General model } & \multicolumn{4}{c}{family universality} & \multicolumn{4}{c}{family universality} \\ \cline{2-5} \cline{6-9} \cline{10-13} \noalign{\smallskip} & $\Lambda^{-}$ & $\Lambda^{+}$ & $\Lambda^{--}$ & $\Lambda^{++}$ & $\Lambda^{-}$ & $\Lambda^{+}$ & $\Lambda^{--}$ & $\Lambda^{++}$ & $\Lambda^{-}$ & $\Lambda^{+}$ & $\Lambda^{--}$ & $\Lambda^{++}$ \\ \hline\hline\hline\noalign{\smallskip} $\eta_{1}$ & \multicolumn{2}{c}{2.1} & \multicolumn{2}{c}{2.3} & \multicolumn{2}{c}{2.7} & \multicolumn{2}{c}{3.0} & \multicolumn{2}{c}{3.1} & \multicolumn{2}{c}{3.5} \\ % $\eta_{APV}$ & 9.8 & 6.0 & 10.4 & 6.6 & 9.7 & 6.1 & 10.3 & 6.7 & 10.9 & 7.5 & 11.7 & 8.4 \\ % $\eta^{ed}_{LL} - \eta^{eu}_{LL}$ & 2.8 & 3.5 & 3.1 & 3.9 & 3.4 & 4.3 & 3.8 & 4.8 & 14.4 & 7.2 & 15.1 & 7.9 \\ \hline\hline\hline\noalign{\smallskip} \end{tabular*} \caption{95\% CL mass-scale limits corresponding to the least constrained coupling combination $\eta_{1}$, atomic parity violating coupling combination $\eta_{APV}$ and $\eta^{ed}_{LL}-\eta^{eu}_{LL}$ combination corresponding to Charged Current contact interaction coupling $\eta^{CC}$ of SU(2) model. As the sign of $\eta_{1}$ is arbitrary, only one value is given.} \label{tab-result3} \end{table*} \begin{figure}[tp] \centerline{\resizebox{\figwidth}{!}{% \includegraphics{zarneckifig5.eps} }} \caption{ Probability functions, calculated within the general contact interaction model, for the two least constrained coupling combinations $\eta_{1}$ and $\eta_{2}$ (upper plots), the atomic parity violating combination $\eta_{APV}$ (lower left) and the CC contact interaction coupling $\eta^{CC}$ induced in the SU(2) model (lower right plot). Note different horizontal scales between upper and lower plots.} \label{fig-prob3} \end{figure} Shown in the same table are mass limits corresponding to the atomic parity violating combination of couplings, \begin{eqnarray} \eta_{APV} & \equiv & \eta^{ed}_{LL} + \eta^{ed}_{LR} - \eta^{ed}_{RL} - \eta^{ed}_{RR} \nonumber \\ & + & \eta^{eu}_{LL} + \eta^{eu}_{LR} - \eta^{eu}_{RL} - \eta^{eu}_{RR} \nonumber \end{eqnarray} $\eta_{APV}$ is close to the most strongly constrained coupling combination (the eigenvector with the highest eigenvalue). Mass scale limits up to about 11~TeV are obtained. The probability function for $\eta_{APV}$ in the case of the general model is included in Figure \ref{fig-prob3}. Also shown in Figure \ref{fig-prob3} is the probability function for the Charged Current contact interaction coupling $\eta^{CC}$ induced in the SU(2) model. The discrepancy between the data and the Standard Model has decreased slightly, as compared with the U1 and U3 SU(2) models. The most probable value of $\eta^{CC}$ is about $2\sigma$ from the Standard Model value, which corresponds to the probability of about 10\%. This discrepancy is observed for the SU(2) model only. When $SU(2)_{L} \times U(1)_{Y}$ universality is not assumed (i.e. in case of the general model and the model with family universality) the corresponding coupling combination is no longer related to the Charged Current sector and is in good agreement with the Standard Model. The corresponding mass scale limits are included in Table \ref{tab-result3}. \section{Predictions} \label{sec-predictions} All presented results are in good agreement with the Standard Model. Nevertheless, an interesting question is whether "new physics" in terms of contact interactions can be expected to show up in high-energy experiments in the near future. The cross-sections corresponding to the "best fit" of the general model (the set of coupling values resulting in the best description of all data, i.e. corresponding to the maximum probability) are compared in Figure \ref{fig-compare} with the HERA, LEP and Tevatron data. In the case of LEP data, the best fit of the general model agrees very well with the Standard Model. The Contact Interaction contribution to the measured cross-section does not exceed 3\% for $\sqrt{s}$ up to 200 GeV. On the other hand, the same model predicts for both HERA and the Tevatron an increase in the cross-section by almost a factor of 2 at the highest $Q^{2}$/$M_{ll}$. In order to verify the significance of these predictions it is unavoidable to consider the statistical uncertainty of these predictions. \begin{figure}[tp] \centerline{\resizebox{!}{\figheight}{% \includegraphics{zarneckifig6.eps} }} \caption{Cross-section deviations from the Standard Model resulting from the general contact interaction model fit (thick solid line) compared with HERA, LEP and the Tevatron data.} \label{fig-compare} \end{figure} Employing the Monte Carlo techniques, the probability function for the contact interaction couplings, ${\cal P}(\vec{\eta})$, is translated into the probability function for relevant cross-section deviations, as described in section \ref{sec-met-prob}. Considered in this analysis are possible deviations from the Standard Model predictions for high-$Q^{2}$ $e^{-}p$ and $e^{+}p$ scattering at HERA\footnote{For the proton beam energy of 920 GeV and the electron/positron beam energy of 27.5 GeV.} (see section \ref{sec-hera}), for the total quark pair production cross-section at LEP (or Next Linear Collider, NLC; see section \ref{sec-lep}) and for the Drell-Yan lepton pair production at the Tevatron (see section \ref{sec-dy}). The probability functions calculated for these processes at selected energy scales are presented in Figure \ref{fig-predhist}. \begin{figure}[tp] \centerline{\resizebox{\figwidth}{!}{% \includegraphics{zarneckifig7.eps} }} \caption{Probability functions for possible deviations from the Standard Model predictions for: $e^{+}p$ and $e^{-}p$ NC DIS cross-section at HERA, at $Q^{2}$ = 30,000 GeV$^{2}$ (upper plots), $e^{+}e^{-}$ total hadronic cross-section at $\sqrt{s}$ = 400 GeV (lower left plot) and Drell-Yan lepton pair production cross-section at the Tevatron, at $M_{ll}$ = 500 GeV (lower right plot).} \label{fig-predhist} \end{figure} The results for HERA, in terms of the 95\% confidence limit bands on the ratio of predicted and the Standard Model cross-sections as a function of Q$^{2}$, are shown in Figures \ref{fig-predhera1} and \ref{fig-predhera2} for the general model and the SU(2) model with family universality, respectively. \begin{figure}[tp] \centerline{\resizebox{\figwidth}{!}{% \includegraphics{zarneckifig8.eps} }} \caption{The 95\% CL limit band on the ratio of predicted to the Standard Model cross-section for $e^{+}p$ and $e^{-}p$ NC DIS scattering at HERA (upper plots) and the 68\% and 95\% CL contours for the possible deviations for scattering of right- and left-handed electrons and positrons at $Q^{2}=$30,000 GeV$^{2}$ (lower plots). The limits are calculated using the general contact interaction model.} \label{fig-predhera1} \end{figure} \begin{figure}[tp] \centerline{\resizebox{\figwidth}{!}{% \includegraphics{zarneckifig9.eps} }} \caption{The 95\% CL limit band on the ratio of predicted to the Standard Model cross-section for $e^{+}p$ and $e^{-}p$ NC DIS scattering at HERA (upper plots) and the 68\% and 95\% CL contours for the possible deviations for scattering of right- and left-handed electrons and positrons at $Q^{2}=$30,000 GeV$^{2}$ (lower plots). The limits are calculated using the SU(2) contact interaction model with family universality.} \label{fig-predhera2} \end{figure} For the $e^{+}p$ NC DIS the uncertainty of these predictions is very big, although the nominal predictions of both models are above the Standard Model. The Standard Model prediction is well within the 95\% confidence level band. For the general model, the increase in the $e^{+}p$ NC DIS cross-section at HERA by up to about 80\% at $Q^{2}$ of 30,000 GeV$^{2}$ would still be consistent with current experimental data. For the SU(2) model the corresponding limit is 63\%. It turns out that the best statistical sensitivity (in single measurement) to possible contact interaction effects is obtained when considering the number of events measured for $Q^{2} \; >$ 15,000 GeV$^{2}$. The allowed increase in the integrated $e^{+}p$ NC DIS cross-section is about 40\% for the general model and about 30\% for the SU(2) model. In order to reach the level of statistical precision, which would allow them to confirm possible discrepancy of this size\footnote{We require that the allowed increase in the cross-section for $Q^{2} >$ 15,000 GeV$^{2}$ (at 95\% CL) should correspond to at least three times the statistical error on the number of events. 5\% systematic uncertainty on the expected number of events is assumed.}, HERA experiments would have to collect $e^{+}p$ luminosities of the order of 100-200 pb$^{-1}$ (depending on the model). This will be possible after the HERA upgrade planned for year 2000. Constraints on the possible deviations from the Standard Model predictions are much stronger in case of $e^{-}p$ NC DIS. This is because the Standard Model cross-section itself is higher, and also because different contact interaction coupling combinations contribute. It is interesting to notice that the possible cross-section increase for $e^{+}p$ NC DIS, which is suggested by global fit results, corresponds to {\bf decrease} in the NC DIS cross-section for $e^{-}p$. For the general model, deviations larger than about 20\% are excluded for $Q^{2} >$15,000 GeV$^{2}$, whereas for the SU(2) model with family universality the limit goes down to about 7\%. When compared with the predicted statistical precision of the future HERA data, this indicates that it will be very hard to detect contact interactions in the future HERA $e^{-}p$ running. For the general model the required luminosity is of the order of 400 pb$^{-1}$. However, the HERA "discovery window" can be visibly enlarged if we consider scattering of polarised electrons and/or positrons. The 68\% and 95\% CL contours for the allowed deviations for scattering of right- and left-handed electrons or positrons are included in Figures \ref{fig-predhera1} (for general model) and \ref{fig-predhera2} (for SU(2) model), at $Q^{2}=$30,000 GeV$^{2}$. In both cases, the cross-section deviations for $e^{+}_{L}p$ and $e^{-}_{R}p$ scattering are less constrained than in case of $e^{+}_{R}p$ and $e^{-}_{L}p$, respectively. For the general model possible deviations for both left- and right-handed projectiles are significantly higher than in the unpolarised case. However, for the SU(2) model, constraints significantly weaker than in the unpolarised case are obtained only for $e^{+}_{L}p$ and $e^{-}_{R}p$ scattering. In both models deviations of up to about 50\% are still allowed for $e^{+}_{L}p$ scattering at $Q^{2}>$15,000 GeV$^{2}$, assuming 60\% polarisation of the positron beam. To observe effects of this size it would be enough to collect luminosity of the order of 70-80 pb$^{-1}$. For $e^{-}_{R}p$ scattering maximum allowed deviations are 28\% and 19\%, for the general and SU(2) models respectively. It means that with 60\% longitudinal $e^{-}_{R}$ polarisation it would be possible to observe significant deviations from Standard Model predictions already for luminosities of the order of 120 pb$^{-1}$ (for the general model). Unfortunately, polarisation can result in significantly higher systematic uncertainties of the Standard Model predictions, which was not considered here. Since the only visible inconsistency between data and the Standard Model is observed in the Charged Current sector (for models assuming $SU(2)_{L} \times U(1)_{Y}$ universality), the interesting question is whether any effect can be observed in high $Q^{2}$ CC DIS at HERA. It turns out that the possible effect is far beyond the HERA sensitivity. The ``best'' $\eta^{CC}$ value (resulting from the SU(2) model fit) corresponds to a {\bf decrease} in the CC DIS cross-section at HERA not greater than 2\% within the accessible $Q^{2}$ range, and a decrease exceeding 5\% is excluded at 95\% CL. At the same confidence level, any increase in the cross-section by a similar amount is excluded. Model predictions for both the total hadronic cross-section at electron-positron collider (LEP or NLC) and Drell-Yan lepton pair production cross-section at the Tevatron are shown in Figures \ref{fig-predlep1} and \ref{fig-predlep2}, for the general contact interaction model and the SU(2) model, respectively. For $e^{-}e^{+} \rightarrow q \bar{q}$ at $\sqrt{s}$ above about 300 GeV upper cross-section limits obtained from both contact interaction models increase rapidly. Cross-section deviations up to a factor of 3 are allowed for $\sqrt{s}\sim$500 GeV. Unfortunately, this energy range will be accessible only in the Next Linear Collider experiment(s). LEP will not go beyond $\sqrt{s}\sim$200 GeV: at this energy the possible deviations from the Standard Model are only about 8\%, which makes possible discovery very difficult. \begin{figure}[tp] \centerline{\resizebox{\figwidth}{!}{% \includegraphics{zarneckifig10.eps} }} \caption{ Left: the 95\% CL limit band on the ratio $\sigma_{SM+CI}/\sigma_{SM}$, where $\sigma_{SM}$ is the Standard Model total hadronic cross-section at LEP/NLC and $\sigma_{SM+CI}$ is the cross-section calculated in the general contact interaction model; Right: the same ratio for the Drell-Yan lepton pair production at the Tevatron. } \label{fig-predlep1} \end{figure} \begin{table}[tp] \begin{center} \begin{tabular}{lcr@{~-}lr@{~-}lr@{~-}l} \hline\hline\hline\noalign{\smallskip} & SM & \multicolumn{6}{c@{}}{Allowed range on 95\% C.L.} \\ \cline{3-8}\noalign{\smallskip} & Value & \multicolumn{2}{c}{General} & \multicolumn{2}{c}{Model with} & \multicolumn{2}{c@{}}{SU(2) model} \\ & (LO) & \multicolumn{2}{c}{model} & \multicolumn{2}{c}{family univ.} & \multicolumn{2}{c@{}}{w. fam. univ.} \\ \hline\hline\hline\noalign{\smallskip} $R_{b}$ & 0.159 & 0.147 & 0.161 & 0.137 & 0.180 & 0.139 & 0.179 \\ % $R_{c}$ & 0.262 & 0.242 & 0.266 & 0.230 & 0.294 & 0.232 & 0.291 \\ % $A^{b}_{FB}$ & 0.601 & & & 0.345 & 0.750 & 0.431 & 0.732 \\ % $A^{c}_{FB}$ & 0.668 & & & 0.469 & 0.750 & 0.551 & 0.738 \\ \hline\hline\hline\noalign{\smallskip} \end{tabular} \end{center} \caption{Leading order Standard Model prediction and the allowed range (on 95\% CL) for the heavy quark production ratios and forward-backward asymmetries, for $e^{+}e^{-}$ annihilation at $\sqrt{s}$=200 GeV, in different contact interaction models.} \label{tab-predict2} \end{table} \begin{figure}[tp] \centerline{\resizebox{\figwidth}{!}{% \includegraphics{zarneckifig11.eps} }} \caption{ Upper left: the 95\% CL limit band on the ratio $\sigma_{SM+CI}/\sigma_{SM}$, where $\sigma_{SM}$ is the Standard Model total hadronic cross-section at LEP/NLC and $\sigma_{SM+CI}$ is the cross-section calculated in the SU(2) contact interaction model; Upper right: the same ratio for the Drell-Yan lepton pair production at the Tevatron; Below: the 68\% and 95\% CL contours for the allowed values of the forward-backward asymmetries and the quark production fractions for $c$ and $b$ quark production at LEP, at $\sqrt{s}$=200 GeV. } \label{fig-predlep2} \end{figure} However, significant deviations from Standard Model predictions are still possible for heavy quark production ratios $R_{c}$ and $R_{b}$, and for the forward-backward asymmetries $A^{c}_{FB}$ and $A^{b}_{FB}$. The 68\% and 95\% CL contours for the values of the forward-backward asymmetry versus the quark production fraction, allowed within the SU(2) model for $c$ and $b$ quark production at $\sqrt{s}$=200 GeV are included in Figure \ref{fig-predlep2}. Allowed ranges for $R_{c}$, $R_{b}$, $A^{c}_{FB}$ and $A^{b}_{FB}$ at $\sqrt{s}$=200 GeV, for different contact interaction models considered, are summarised in Table \ref{tab-predict2}. For the general model, in which contact interactions are limited to the first quark generation only, variations of $R_{b}$ and $R_{c}$ are still possible, due to the possible changes in $u\bar{u}$ and $d\bar{d}$ production cross-sections. However, parton level forward-backward asymmetries $A^{c}_{FB}$ and $A^{b}_{FB}$ do not depend on contact interaction couplings in this model. Therefore limits for $A^{c}_{FB}$ and $A^{b}_{FB}$ are not reported for the general model. Heavy quark observables considered here are least constrained for the model with family universality. Large effects are still possible in this model for both production fractions and asymmetries. Deviations up to about 13\% are possible for $R_{b}$ and $R_{c}$. Least constrained by the existing experimental data is the forward-backward asymmetry for the $b\bar{b}$ production $A^{b}_{FB}$, where deviations from the Standard Model prediction by up to 40\% are still possible. Note that for the general model $R_{b}$ and $R_{c}$ are 100\% correlated, whereas for models with the family universality they are 100\% anti-correlated. It seems that the best place to study contact interactions in the nearest future is the Tevatron, which should run again after being upgraded in the year 2000. If there is any "new physics" corresponding to the contact interaction model it is very likely to show up in Drell-Yan lepton pair production for masses above 200-300 GeV. Moreover, upper limits on possible deviations from the Standard Model predictions are much higher than in case of HERA and LEP/NLC. For $M_{ll}=$500 GeV, which should be easily accessible with increased luminosity, cross-section deviations up to a factor of 5 are still not excluded. Upper limits on the cross-section deviations from the Standard Model predictions, derived on 95\% confidence level in different contact interaction models are summarised in Table~\ref{tab-predict}. \begin{table}[tp] \begin{center} \begin{tabular}{llrrr} \hline\hline\hline\noalign{\smallskip} & & \multicolumn{3}{c}{Limits on $\Delta \sigma / \sigma_{SM}$ [\%]} \\ \cline{3-5}\noalign{\smallskip} Reaction & Energy & General & Model with & SU(2) model \\ & scale & model & family univ. & w. family univ. \\ \hline\hline\hline\noalign{\smallskip} $e^{+}p$ NC DIS & $Q^{2}$=10000 GeV$^{2}$ & 11 & 10 & 9 \\ $\sqrt{s}=$318 GeV & $Q^{2}$=20000 GeV$^{2}$ & 36 & 30 & 28 \\ & $Q^{2}$=30000 GeV$^{2}$ & 81 & 65 & 63 \\ & $Q^{2}$=50000 GeV$^{2}$ & 220 & 180 & 170 \\ \hline\noalign{\smallskip} % $e^{-}p$ NC DIS & $Q^{2}$=10000 GeV$^{2}$ & 8 & 4 & 3 \\ $\sqrt{s}=$318 GeV & $Q^{2}$=20000 GeV$^{2}$ & 18 & 8 & 7 \\ & $Q^{2}$=30000 GeV$^{2}$ & 28 & 13 & 11 \\ & $Q^{2}$=50000 GeV$^{2}$ & 49 & 26 & 21 \\ \hline\noalign{\smallskip} % $e^{-}e^{+} \rightarrow q \bar{q}$ & $\sqrt{s}$=175 GeV & 5 & 5 & 6 \\ & $\sqrt{s}$=200 GeV & 8 & 8 & 8 \\ & $\sqrt{s}$=225 GeV & 14 & 13 & 11 \\ & $\sqrt{s}$=250 GeV & 26 & 24 & 16 \\ & $\sqrt{s}$=300 GeV & 65 & 61 & 35 \\ & $\sqrt{s}$=400 GeV & 185 & 185 & 110 \\ \hline\noalign{\smallskip} % $p \bar{p} \rightarrow l^{+} l^{-} X $ & $M_{ll}$=200 GeV & 17 & 12 & 12 \\ $\sqrt{s} = $ 1800 GeV & $M_{ll}$=300 GeV & 64 & 55 & 38 \\ & $M_{ll}$=400 GeV & 190 & 185 & 95 \\ & $M_{ll}$=500 GeV & 440 & 450 & 210 \\ \hline\hline\hline\noalign{\smallskip} \end{tabular} \end{center} \caption{Upper limits (on 95\% CL) on cross-section deviations from the Standard Model predictions in different contact interaction models. Considered are $e^{+}p$ and $e^{-}p$ scattering at HERA, total hadronic cross-section at LEP/NLC and Drell-Yan lepton pair production at the Tevatron, as indicated in the table.} \label{tab-predict} \end{table} \begin{figure}[tp] \centerline{\resizebox{!}{\figheight}{% \includegraphics{zarneckifig12.eps} }} \caption{The 68\% and 95\% CL contours for the possible deviation from the Standard Model predictions, for different combinations of measurements. Considered are: $e^{+}p$ and $e^{-}p$ NC DIS cross-section at HERA, at $Q^{2}$ = 30,000 GeV$^{2}$, total $e^{+}e^{-} \rightarrow q\bar{q}$ cross-section at $\sqrt{s}$ = 400 GeV and Drell-Yan lepton pair production cross-section at the Tevatron, at $M_{ll}$ = 500 GeV, as indicated on the plot. The limits are calculated using the general contact interaction model.} \label{fig-pred2d1} \end{figure} \begin{figure}[tp] \centerline{\resizebox{!}{\figheight}{% \includegraphics{zarneckifig13.eps} }} \caption{The 68\% and 95\% CL contours for the possible deviation from the Standard Model predictions, for different combinations of measurements. Considered are: $e^{+}p$ and $e^{-}p$ NC DIS cross-section at HERA, at $Q^{2}$ = 30,000 GeV$^{2}$, total $e^{+}e^{-} \rightarrow q\bar{q}$ cross-section at $\sqrt{s}$ = 400 GeV and Drell-Yan lepton pair production cross-section at the Tevatron, at $M_{ll}$ = 500 GeV, as indicated on the plot. The limits are calculated using the SU(2) contact interaction model with family universality.} \label{fig-pred2d2} \end{figure} When considering possible future discoveries at high-energy experiments, it is also interesting to study the relation between effects observed at different experiments. The 68\% and 95\% CL contours for the sizes of the allowed deviation from the Standard Model predictions, for different measurement combinations, are shown in Figures \ref{fig-pred2d1} and \ref{fig-pred2d2}, for the general contact interaction model and for the SU(2) model with family universality, respectively. In both cases, clear correlation is observed between the Drell-Yan cross-section deviation at the Tevatron and the hadronic $e^{+}e^{-}$ cross-section at LEP/NLC. Possible cross-section increase at the Tevatron has to be accompanied by the increase in the hadronic cross-section at LEP/NLC. Similar correlation is observed between the hadronic $e^{+}e^{-}$ cross-section at LEP/NLC and $e^{+}p$ NC DIS cross-section at HERA for SU(2) model. Another interesting observation is that the possible decrease in the $e^{-}p$ NC DIS cross-section at HERA should be related to the increase in both Tevatron and LEP/NLC cross-section. In other cases correlations between different measurements are weak. This shows that contact interaction searches at LEP, the Tevatron and HERA are, to large extent, independent. Data from all types of experiments are necessary to constraint contact interaction model in general case. \clearpage \section{Summary} \label{sec-summary} Data from HERA, LEP, the Tevatron and low energy experiments were used to constrain electron-quark contact interactions. The contact interaction mass scale limits obtained for different one-parameter models range from 5.1 to about 18 TeV. Using the most general approach, in which all couplings to are allowed to vary independently, any contact interactions with mass scale below 2.1 TeV are excluded at 95\% CL. This limit can be raised to 3.1~TeV by assuming $SU(2)_{L} \times U(1)_{Y}$ and quark/lepton family universality. There is a slight hint on possible ``new physics'' in the Charged Current sector (related to Neutral Current contact interactions by $SU(2)_{L} \times U(1)_{Y}$ universality), where the discrepancy between data and the Standard Model is at the $2\sigma$ level. The mass scale of new Charged Current interactions suggested by the data is of the order of 10~TeV. However, this effect - if real - would have negligible impact on predictions for future collider results. The limits on possible effects to be observed in future HERA, LEP and Tevatron running are estimated. Possible deviations from the Standard-Model predictions for total hadronic cross-section at LEP and $e^{-}p$ scattering cross-section at HERA, are already strongly limited by existing data. However, improved experimental sensitivity to new interactions should result from the measurement of heavy quark production ratios and asymmetries at LEP, as well as from polarised electron scattering at HERA. Sizable effects are still not excluded for $e^{+}p$ NC DIS at HERA and the required statistical precision of the data should be accessible after HERA upgrade. The best "discovery potential" seems to come from future Tevatron running, where significant deviations from the Standard Model predictions are still allowed. For Drell-Yan lepton pair production cross-section deviations at $M_{ll}$=500 GeV up to a factor of 5 are still not excluded. However, all experiments should continue to analyse their data in terms of possible new electron-quark interactions, as constraints resulting from different experiments are, to large extent, complementary. \vfill \section*{Acknowledgements} I would like to thank all members of the Warsaw HEP group and of the ZEUS Collaboration for support, encouragement, many useful comments and suggestions. Special thanks are due to U.Katz for very productive discussions and many valuable comments to this paper, and to Prof. A.K.Wr\'{o}blewski for his help in preparing the final version of it. I am also grateful to M.Lancaster and W.Sakumoto from CDF, A.Gupta and A.Kotwal from D0, I.Tomalin from {\sc Aleph}, A.Olchevski from {\sc Delphi}, F.Filthaut from L3 and P.Ward from {\sc Opal} for their assistance in gathering the relevant experimental data and in understanding details of the measurements. This work has been partially supported by the Polish State Committee for Scientific Research (grant No. 2 P03B 035 17). \clearpage
\section{INTRODUCTION} The best determined neutron star masses are found in binary pulsars and all lie in the range $1.35\pm 0.04 M_\odot$ (see Thorsett and Chakrabarty 1999) except for the nonrelativistic pulsar PSR J1012+5307 of mass\footnote{95\% conf. limits or $\sim2\sigma$} $M=(2.1\pm 0.8)M_\odot$ (van Paradijs 1998). Several X-ray binary masses have been measured of which the heaviest are Vela X-1 with $M=(1.9\pm 0.2)M_\odot$ (Barziv et al., 1999) and Cygnus X-2 with $M=(1.8\pm 0.4)M_\odot$ (Orosz \& Kuulkers 1999). The recent discovery of high-frequency brightness oscillations in low-mass X-ray binaries provides a promising new method for determining masses and radii of neutron stars (see Miller, Lamb, \& Psaltis 1998). The kilohertz quasi-periodic oscillations (QPO) occur in pairs and are most likely the orbital frequencies $\nu_{QPO}=(1/2\pi)\sqrt{GM/R_{orb}^3}$ of accreting matter in Keplerian orbits around neutron stars of mass $M$ and its beat frequency with the neutron star spin, $\nu_{QPO}-\nu_s$. According to Zhang, Strohmayer, \& Swank 1997, Kaaret, Ford, \& Chen (1997) the accretion can for a few QPO's be tracked to its innermost stable orbit, $R_{ms}=6GM/c^2$. For slowly rotating stars the resulting mass is $M\simeq2.2M_\odot({\mathrm{kHz}}/\nu_{QPO})$. For example, the maximum frequency of 1060 Hz upper QPO observed in 4U 1820-30 gives $M\simeq 2.25M_\odot$ after correcting for the $\nu_s\simeq275$ Hz neutron star rotation frequency. If the maximum QPO frequencies of 4U 1608-52 ($\nu_{QPO}=1125$ Hz) and 4U 1636-536 ($\nu_{QPO}=1228$ Hz) also correspond to innermost stable orbits the corresponding masses are $2.1M_\odot$ and $1.9M_\odot$. Such large masses severely restrict the equation of state (EoS) for dense matter as addressed in the following. Recent models for the nucleon-nucleon interaction have reduced the uncertainty in the nuclear EoS allowing for more reliable calculations of neutron star properties, see Akmal, Pandharipande, \& Ravenhall (1998) and Engvik et al.~(1997). Likewise, recent realistic effective interactions for nuclear matter obeying causality at high densities, constrain the EoS severely and thus also the maximum masses of neutron stars, see Akmal, Pandharipande, \& Ravenhall (1998) and Kalogera \& Baym (1996). We will here elaborate on these analyses by incorporating causality smoothly in the EoS for nuclear matter and allow for first and second order phase transitions to, e.g., quark matter. Finally, results are compared with observed neutron star masses and concluding remarks are made. \section{THE NUCLEAR EQUATION OF STATE} For the discussion of the gross properties of neutron stars we will use the optimal EoS of Akmal, Pandharipande, \& Ravenhall (1998) (specifically the Argonne $V18 + \delta v +$ UIX$^*$ model- hereafter APR98), which is based on the most recent models for the nucleon-nucleon interaction with the inclusion of a parametrized three-body force and relativistic boost corrections. The EoS for nuclear matter is thus known to some accuracy for densities up to a few times nuclear saturation density $n_0=0.16$ fm$^{-3}$. We parametrize the APR98 EoS by a simple form for the compressional and symmetry energies that gives a good fit around nuclear saturation densities and smoothly incorporates causality at high densities such that the sound speed approaches the speed of light. This requires that the compressional part of the energy per nucleon is quadratic in nuclear density with a minimum at saturation but linear at high densities \begin{eqnarray} {\cal E} &=& E_{comp}(n) + S(n)(1-2x)^2 \nonumber\\ &=& {\cal E}_0 u\frac{u-2-s}{1+s u} +S_0 u^\gamma (1-2x)^2. \label{eq:EA} \end{eqnarray} Here, $n=n_p+n_n$ is the total baryon density, $x=n_p/n$ the proton fraction and $u=n/n_0$ is the ratio of the baryon density to nuclear saturation density. The compressional term is in Eq.\ (\ref{eq:EA}) parametrized by a simple form which reproduces the saturation density and the binding energy per nucleon ${\cal E}_0=15.8$MeV at $n_0$ of APR98. The ``softness'' parameter $s\simeq 0.2$, which gave the best fit to the data of APR98 (see Heiselberg \& Hjorth-Jensen 1999) is determined by fitting the energy per nucleon of APR98 up to densities of $n\sim 4n_0$. For the symmetry energy term we obtain $S_0=32$ MeV and $\gamma=0.6$ for the best fit. The proton fraction is given by $\beta$-equilibrium at a given density. The one unknown parameter $s$ expresses the uncertainty in the EoS at high density and we shall vary this parameter within the allowed limits in the following with and without phase transitions to calculate mass, radius and density relations for neutron stars. The ``softness'' parameter $s$ is related to the incompressibility of nuclear matter as $K_0=18{\cal E}_0/(1+s)\simeq 200$MeV. It agrees with the poorly known experimental value (Blaizot, Berger, Decharge, \& Girod 1995), $K_0\simeq 180-250$MeV which does not restrict it very well. From $(v_s/c)^2=\partial P/\partial (n\cal{E})$, where $P$ is the pressure, and the EoS of Eq.\ (\ref{eq:EA}), the causality condition $v_s\le c$ requires \begin{equation} s \ga \sqrt{\frac{{\cal E}_0}{m_n}} \simeq 0.13 \,,\label{causal} \end{equation} where $m_n$ is the mass of the nucleon. With this condition we have a causal EoS that reproduces the data of APR98 at densities up to $0.6\sim 0.7$ fm$^{-3}$. In contrast, the EoS of APR98 becomes superluminal at $n\approx 1.1$ fm$^{-3}$. For larger $s$ values the EoS is softer which eventually leads to smaller maximum masses of neutron stars. The observed $M\simeq 1.4M_\odot$ in binary pulsars restricts $s$ to be less than $0.4-0.5$ depending on rotation as shown in calculations of neutron stars below. In Fig.\ \ref{fig1} we plot the sound speed $(v_s/c)^2$ for various values of $s$ and that resulting from the microscopic calculation of APR98 for $\beta$-stable $pn$-matter. The form of Eq.~(\ref{eq:EA}), with the inclusion of the parameter $s$, provides a smooth extrapolation from small to large densities such that the sound speed $v_s$ approaches the speed of light. For $s=0.0$ ($s=0.1$) the EoS becomes superluminal at densities of the order of 1 (6) fm$^{-3}$. The sound speed of Kalogera \& Baym (1996) is also plotted in Fig.\ \ref{fig1}. It jumps discontinuously to the speed of light at a chosen density. With this prescription they were able to obtain an optimum upper bound for neutron star masses and obey causality. This prescription was also employed by APR98, see Rhoades \& Ruffini (1974) for further details. The EoS is thus discontinuously stiffened by taking $v_s=c$ at densities above a certain value $n_c$ which, however, is lower than $n_{s}=5n_0$ where their nuclear EoS becomes superluminal. This approach stiffens the nuclear EoS for densities $n_c<n<n_s$ but softens it at higher densities. Their resulting maximum masses lie in the range $2.2M_\odot\la M\la 2.9M_\odot$. Our approach however, incorporates causality by reducing the sound speed smoothly towards the speed of light at high densities. Therefore our approach will not yield an absolute upper bound on the maximum mass of a neutron star but gives reasonable estimates based on modern EoS around nuclear matter densities, causality constraints at high densities and a smooth extrapolation between these two limits (see Fig. 1). \section{PHASE TRANSITIONS} The physical state of matter in the interiors of neutron stars at densities above a few times normal nuclear matter densities is essentially unknown and many first and second order phase transitions have been speculated upon. We will specifically study the hadron to quark matter transition at high densities, but note that other transitions as, e.g., kaon and/or pion condensation or the presence of other baryons like hyperons also soften the EoS and thus further aggravate the resulting reduction in maximum masses. Hyperons appear at densities typically of the order $2 n_0$ and result in a considerable softening of the EoS, see e.g., Balberg, Lichenstadt, \& Cook (1998). Typically, most equations of state with hyperons yield masses around $1.4-1.6 M_{\odot}$. Here however, in order to focus on the role played by phase transitions in neutron star matter, we will assume that a phase transition from nucleonic to quark matter takes place at a certain density. We will for simplicity employ the bag model in our actual studies of quark phases and neutron star properties. In the bag model the quarks are assumed to be confined to a finite region of space, the so-called 'bag', by a vacuum pressure $B$. Adding the Fermi pressure and interactions computed to order $\alpha_s=g^2/4\pi$, where $g$ is the QCD coupling constant, the total pressure for 3 massless quarks of flavor $f=u,d,s$, is (see Kapusta (1988), \begin{equation} P=\frac{3\mu_f^4}{4\pi^2}(1-\frac{2}{\pi}\alpha_s) -B +P_e+P_\mu \,, \label{pquark} \end{equation} where $P_{e,\mu}$ are the electron and muon pressure, e.g., $P_e=\mu_e^4/12\pi^2$. A Fermi gas of quarks of flavor {\em i} has density $n_i = k_{Fi}^3/\pi^2$, due to the three color states. The value of the bag constant {\em B} is poorly known, and we present results using two representative values, $B=150$ MeVfm$^{-3}$ and $B=200$ MeVfm$^{-3}$. We take $\alpha_s=0.4$. However, similar results can be obtained with smaller $\alpha_s$ and larger $B$ (Madsen 1998). The quark and nuclear matter mixed phase described in Glendenning (1992) has continuous pressures and densities due to the general Gibbs criteria for two-component systems. There are no first order phase transitions but at most two second order phase transitions. Namely, at a lower density, where quark matter first appears in nuclear matter, and at a very high density (if gravitationally stable), where all nucleons are finally dissolved into quark matter. This mixed phase does, however, not include local surface and Coulomb energies of the quark and nuclear matter structures. If the interface tension between quark and nuclear matter is too large, the mixed phase is not favored energetically due to surface and Coulomb energies associated with forming these structures (Heiselberg, Pethick, \& Staubo 1993). The neutron star will then have a core of pure quark matter with a mantle of nuclear matter surrounding it and the two phases are coexisting by a first order phase transition or Maxwell construction, see Fig. 2. For a small or moderate interface tension the quarks are confined in droplet, rod- and plate-like structures as found in the inner crust of neutron stars (Lorenz, Ravenhall, \& Pethick 1993). \section{NEUTRON STAR PROPERTIES} In order to obtain the mass and radius of a neutron star, we have solved the Tolman-Oppenheimer-Volkov equation with and without rotational corrections following the approach of Hartle (1967). The equations of state employed are given by the $pn$-matter EoS with $s =0.13, 0.2, 0.3, 0.4$ with nucleonic degrees of freedon only. In addition we have selected two representative values for the bag-model parameter $B$, namely 150 and 200 MeVfm$^{-3}$ for our discussion on eventual phase transitions. The quark phase is linked with our $pn$-matter EoS from Eq.\ (\ref{eq:EA}) with $s=0.2$ through either a mixed phase construction or a Maxwell construction, see Heiselberg and Hjorth-Jensen (1999) for further details. For $B=150$ MeVfm$^{-3}$, the mixed phase begins at 0.51 fm$^{-3}$ and the pure quark matter phase begins at $1.89$ fm$^{-3}$. Finally, for $B=200$ MeVfm$^{-3}$, the mixed phase starts at $0.72$ fm$^{-3}$ while the pure quark phase starts at $2.11$ fm$^{-3}$. In case of a Maxwell construction, in order to link the $pn$ and the quark matter EoS, we obtain for $B=150$ MeVfm$^{-3}$ that the pure $pn$ phase ends at $0.92$ fm$^{-3}$ and that the pure quark phase starts at $1.215$ fm$^{-3}$, while the corresponding numbers for $B=200$ MeVfm$^{-3}$ are $1.04$ and $1.57$ fm$^{-3}$. As can be seen from Fig.\ \ref{fig2} none of the equations of state from either the pure $pn$ phase or with a mixed phase or Maxwell construction with quark degrees of freedom, result in stable configurations for densities above $\sim 10 n_0$, implying thereby that none of the stars have cores with a pure quark phase. The EoS with $pn$ degrees of freedom have masses $M\la2.2M_{\odot}$ when rotational corrections are accounted for. With the inclusion of the mixed phase, the total mass is reduced since the EoS is softer. For pure quark stars there is only one energy scale namely $B$ which provides a homology transformation (Madsen 1998) and the maximum mass is $M_{max}=2.0M_\odot (58{\rm MeV fm^{-3}}/B)^{1/2}$ (for $\alpha_s=0$). However, for $B\ga 58{\rm MeV fm^{-3}}$ a nuclear matter mantle has to be added and for $B\la 58{\rm MeV fm^{-3}}$ quark matter has lower energy per baryon than $^{56}$Fe and is thus the ground state of strongly interacting matter. Unless the latter is the case, we can thus exclude the existence of $2.2-2.3M_\odot$ quark stars. In Fig.\ \ref{fig3} we show the mass-radius relations for the various equations of state. The shaded area represents the allowed masses and radii for $\nu_{QPO}=1060$ Hz of 4U 1820-30. Generally, \begin{eqnarray} 2GM < R < \left(\frac{GM}{4\pi^2\nu_{QPO}^2}\right)^{1/3} \,, \end{eqnarray} where the lower limit ensures that the star is not a black hole, and the upper limit that the accreting matter orbits outside the star, $R<R_{orb}$. Furthermore, for the matter to be outside the innermost stable orbit, $R>R_{ms}=6GM$, requires that \begin{eqnarray} M &\la& \frac{1+0.75j}{12\sqrt{6}\pi G\nu_{QPO}} \label{Mms} \\ &\simeq& 2.2 M_\odot (1+0.75j)\frac{{\rm kHz}}{\nu_{QPO}} \,, \nonumber \end{eqnarray} where $j=2\pi c\nu_sI/M^2$ is a dimensionless measure of the angular momentum of the star with moment of inertia $I$. The upper limit in Eq. (\ref{Mms}) is the mass when $\nu_{QPO}$ corresponds to the innermost stable orbit. According to Zhang, Smale, Strohmayer \& Swank (1998) this is the case for 4U 1820-30 since $\nu_{QPO}$ saturates at $\sim1060$~Hz with increasing count rate. The corresponding neutron star mass is $M\sim 2.2-2.3M_\odot$ which leads to several interesting conclusions as seen in Fig.\ \ref{fig3}. Firstly, the stiffest EoS allowed by causality ($s\simeq 0.13-0.2$) is needed. Secondly, rotation must be included which increases the maximum mass and corresponding radii by 10-15\% for $\nu_s\sim 300$~Hz. Thirdly, a phase transition to quark matter below densities of order $\sim 5 n_0$ can be excluded, corresponding to a restriction on the bag constant $B\ga200$ MeVfm$^{-3}$. These maximum masses are smaller than those of APR98 and Kalogera \& Baym (1996) who, as discussed above, obtain upper bounds on the mass of neutron stars by discontinuously setting the sound speed to equal the speed of light above a certain density, $n_c$. By varying the density $n_c=2\to 5n_0$ the maximum mass drops from $2.9\to 2.2M_\odot$. In our case, incorporating causality smoothly by introducting the parameter $s$ in Eq.\ (\ref{eq:EA}), the EoS is softened at higher densities in order to obey causality, and yields a maximum mass which instead is slightly lower than the $2.2M_\odot$ derived in APR98 for nonrotating stars. If the QPOs are not from the innermost stable orbits and one finds that even accreting neutron stars have small masses, say like the binary pulsars, $M\la1.4M_\odot$, this may indicate that heavier neutron stars are not stable. Therefore, the EoS is soft at high densities $s\ga0.4$ or that a phase transition occurs at few times nuclear matter densities. For the nuclear to quark matter transition this would require $B<80$ MeVfm$^{-3}$ for $s=0.2$. For such small bag parameters there is an appreciable quark and nuclear matter mixed phase in the neutron star interior but even in these extreme cases a pure quark matter core is not obtained for stable neutron star configurations. A third QPO frequency referred to as Horisontal Branch Oscillations around $\nu_{HBO}\simeq 20-50$ Hz has been suggested to be caused by Lense-Thirring precession at the inner border of the accretion disk (Stella \& Vietri 1998) \begin{eqnarray} \nu_{LT} &=& \frac{8\pi^2 I}{c^2M} \nu_s\nu^2_{QPO} \nonumber\\ &\simeq& \frac{13.2}{50{\rm km^2}}\frac{I}{M} \frac{\nu_s}{300{\rm H}z}\left(\frac{\nu_{QPO}}{{\rm 1kHz}}\right)^2 \,. \label{LT} \end{eqnarray} However, even for the stiffest EoS $s\simeq 0.13-0.2$ we calculate moment of inertia and Lense-Thirring frequencies from Eq.~(\ref{LT}), which are a factor $\sim4$ below the observed $\nu_{HBO}$ thus confirming analyses of Schaab \& Weigel (1999), Psaltis et al.~(1999) and Kalogera \& Psaltis (1999). \section{SUMMARY} Modern nucleon-nucleon potentials have reduced the uncertainties in the calculated EoS. Using the most recent realistic effective interactions for nuclear matter of APR98 with a smooth extrapolation to high densities including causality, the equation of state could be constrained by a ``softness'' parameter $s$ which parametrizes the unknown stiffness of the EoS at high densities. Maximum masses were calculated for rotating neutron stars with and without first and second order phase transitions to, e.g., quark matter at high densities. The calculated bounds for maximum masses leaves two natural options when compared to the observed neutron star masses: \begin{itemize} \item {\bf Case I}: {\it The large masses of the neutron stars in QPO 4U 1820-30 ($M=2.3M_\odot$), PSR J1012+5307 ($M=2.1\pm0.4 M_\odot$), Vela X-1 ($M=1.9\pm0.1 M_\odot$), and Cygnus X-2 ($M=1.8\pm0.4 M_\odot$), are confirmed and complemented by other neutron stars with masses around $\sim 2M_\odot$.} As a consequence, the EoS of dense nuclear matter is severely restricted and only the stiffest EoS consistent with causality are allowed, i.e., softness parameter $0.13\le s\la0.2$. Furthermore, any significant phase transition at densities below $< 5n_0$ can be excluded. That the radio binary pulsars all have masses around $1.4M_\odot$ is then probably due to the formation mechanism in supernovae where the Chandrasekhar mass for iron cores are $\sim1.5M_\odot$. Neutron stars in binaries can subsequently acquire larger masses by accretion as X-ray binaries. \item {\bf Case II}: {\it The heavy neutron stars prove erroneous by more detailed observations and only masses like those of binary pulsars are found.} If accretion does not produce neutron stars heavier than $\ga1.4M_\odot$, this indicates that heavier neutron stars simply are not stable which in turn implies a soft EoS, either $s> 0.4$ or a significant phase transition must occur already at a few times nuclear saturation densities. \end{itemize}
\section{Introduction} Antiproton annihilation on nuclei provides new possibilities for studying open charm production, exploring the properties of charmed particles in nuclear matter and measuring the interaction of charmed hadrons. Hatsuda and Kunihiro~\cite{Hatsuda} proposed that the light quark condensates may be substantially reduced in hot and dense matter and that as a result hadron masses would be modified. At low density the ratio of the scalar hadron mass in medium to that in vacuum can be directly linked to the ratio of the quark condensates~\cite{Nelson,Brown,Hatsuda1,Lutz,Saito1}. Even if the change in the ratio of the quark condensates is small, the absolute difference between the in-medium and vacuum masses of the hadron is expected to be larger for the heavy hadrons. In practice, any detection of the modification of the mass of a hadron in matter deals with the measurement of effect associated with this absolute difference. It was found in Refs.~\cite{Klingl,Hayahigaki} that the in medium change of quark condensates is smaller for heavier quarks, $s$ and $c$, than those for the light quarks, $u$ and $d$. Thus the in-medium modification of the properties of heavy hadrons may be regarded as being controlled mainly by the light quark condensates. \, As a consequence we \, expect that charmed mesons, which consist of a light quark and heavy $c$ quark, should serve as suitable probes of the in-medium modification of hadron properties. As for the $\bar{K}$ and $K$-mesons, with their quark contents $\bar{q}s$ and $q\bar{s}$ ($q{=}u,d$ light quarks), respectively, the $D$ ($\bar{q}c$) and $\bar{D}$ ($q\bar{c}$) mesons will satisfy different dispersion relations in nuclear matter because of the different sign of the $q$ and $\bar{q}$ vector coupling. Some experimental confirmation of this effect has come from measurements~\cite{Kaos,Schroter,Ritman,Barth,Shin,Li,Cassing,Li1,Cassing1} of $K^-$ and $K^+$-meson production from heavy ion collisions. The $D^+$ and $D^-$ production from $\bar{p}A$ annihilation might yield an even cleaner signal for the in-medium modification of the $D$ and $\bar{D}$ masses. Because of charm conservation, $D$ and $\bar{D}$ mesons are produced pairwise and can be detected by their semileptonic decay channels. The threshold for the $\bar{p}N{\to}D\bar{D}$ reaction in vacuum opens at an antiproton energy around 5.57~GeV, but it is lowered in the $\bar{p}A$ annihilation by the in-medium modification of the $D$ and $\bar{D}$ masses as well as by Fermi motion. The interaction of the $D$-mesons with nuclear matter is of special interest~\cite{Kharzeev}. Note that the $DN$ interaction should be very different from that of charmonia ($J{/}\Psi{N}$), since the interaction between the nucleons and the heavy mesons which do not contain $u$ and $d$ quarks is expected to proceed predominantly through gluon exchange. On the other hand, as for $\bar{K}$-mesons, the $D$-mesons might be strongly absorbed in matter because of the charm exchange reaction $DN{\to}\Lambda_c\pi$, while the $\bar{D}$-mesons should not be absorbed. As will be shown later, the specific conditions of the $D\bar{D}$ pair production in $\bar{p}A$ annihilation provide a very clean and almost model independent opportunity for the experimental reconstruction of the charm-exchange mechanism. \section{D-meson mass in nuclei} As far as the meson properties in free space are concerned, the Bethe-Salpeter (BS) and Dyson-Schwinger (DS) approaches have been widely used~\cite{mir}. The application of BS approach to the description~\cite{Weiss} of heavy and light quarks system allows well to describe the D and B meson properties in free space. The DS approach at finite baryon density was used~\cite{Blaschke} for the calculation of the in-medium properties of $\rho$, $\omega$ and $\phi$ mesons. The modification of the $\rho$ and $\omega$ meson masses resulting from DS equation is close to the calculations with the quark-meson coupling (QMC) model~\cite{qmc}, while the $\phi$-meson mass reduction from Ref.~\cite{Blaschke} is larger as compared to QMC. However, these approaches are not still developed well for the system of finite baryon density and we could not compare their results for the hadron properties in nuclear matter with the predictions from other models. Here, we use the quark-meson coupling model~\cite{qmc}, which has been successfully applied not only to the problems of conventional nuclear physics~\cite{Guichon,Saito2} but also to the studies of meson properties in a nuclear medium~\cite{Tsushima1,Tsushima2,tony}. Furthermore, the properties of the $D$ meson (also $B$) in free space are well described in an MIT bag model~\cite{MIT}. A detailed description of the Lagrangian density and the mean-field equations of motion needed to describe a finite nucleus are given in Refs.~\cite{Guichon,Saito2}. At position \mbox{\boldmath$r$} in a nucleus (the coordinate origin is taken at the center of the nucleus), the Dirac equations for the quarks and antiquarks in the $D$ and $\bar{D}$ meson bags, neglecting the Coulomb force, are given by~\cite{Tsushima1}: \begin{eqnarray} \left[ i \gamma \cdot \partial_x - (m_q - V^q_\sigma(\mbox{\boldmath $r$})) \mp \gamma^0 \left( V^q_\omega(\mbox{\boldmath $r$}) + \frac{1}{2} V^q_\rho(\mbox{\boldmath $r$}) \right) \right] \nonumber \\ \times \left( \begin{array}{c} \psi_u(x) \\ \psi_{\bar{u}}(x) \\ \end{array} \right) = 0, \label{diracu} \\ \left[ i \gamma \cdot \partial_x - (m_q - V^q_\sigma(\mbox{\boldmath $r$})) \mp \gamma^0 \left( V^q_\omega(\mbox{\boldmath $r$}) - \frac{1}{2} V^q_\rho(\mbox{\boldmath $r$}) \right) \right] \nonumber \\ \times \left(\begin{array}{c} \psi_d(x) \\ \psi_{\bar{d}}(x) \\ \end{array} \right) = 0, \label{diracd} \\ \left[ i \gamma \cdot \partial_x - m_{c} \right] \psi_{c} (x)\,\, ({\rm or}\,\, \psi_{\bar{c}}(x)) = 0. \label{diracsc} \end{eqnarray} The mean-field potentials for a bag centered at position \mbox{\boldmath $r$} in the nucleus are defined by $V^q_\sigma(\mbox{\boldmath $r$}){=}g^q_\sigma \sigma(\mbox{\boldmath $r$})$, $V^q_\omega(\mbox{\boldmath $r$}){=}$ $g^q_\omega \omega(\mbox{\boldmath $r$})$ and $V^q_\rho(\mbox{\boldmath $r$}){=}g^q_\rho b(\mbox{\boldmath $r$})$, with $g^q_\sigma$, $g^q_\omega$ and $g^q_\rho$ the corresponding quark and meson-field coupling constants. (Note that we have neglected the small variation of the scalar and vector mean-fields inside the meson bag due to its finite size~\cite{Guichon}.) The mean meson fields are calculated self-consistently by solving Eqs.~(23) -- (30) of Ref.~\cite{Saito2}, namely, by solving a set of coupled non-linear differential equations for static, spherically symmetric nuclei, resulting from the variation of the effective Lagrangian density involving the quark degrees of freedom and the scalar, vector and Coulomb fields in mean field approximation. The normalized, static solution for the ground state quarks or antiquarks in the meson bags may be written as: \begin{eqnarray} \psi_f (x) = N_f e^{- i \epsilon_f t / R_j^*} \psi_f (\mbox{\boldmath $x$}), \qquad (j = D, \bar{D}), \label{wavefunction} \end{eqnarray} where $f{=}u$, $\bar{u}$, $d$, $\bar{d}$, $c$, $\bar{c}$ refers to quark flavors, and $N_f$ and $\psi_f(\mbox{\boldmath $x$})$ are the normalization factor and corresponding spin and spatial part of the wave function. The bag radius in medium, $R_j^*$, which generally depends on the hadron species to which the quarks and antiquarks belong, will be determined through the stability condition for the (in-medium) mass of the meson against the variation of the bag radius~\cite{Guichon,Saito2,Tsushima1} (see also Eq.~(\ref{equil})). The eigenenergies $\epsilon_f$ in Eq.~(\ref{wavefunction}) in units of $1/R_j^*$ are given by \begin{eqnarray} \left( \begin{array}{c} \epsilon_u(\mbox{\boldmath $r$}) \\ \epsilon_{\bar{u}}(\mbox{\boldmath $r$}) \end{array} \right) &=& \Omega_q^*(\mbox{\boldmath $r$}) \pm R_j^* \left( V^q_\omega(\mbox{\boldmath $r$}) + \frac{1}{2} V^q_\rho(\mbox{\boldmath $r$}) \right), \label{uenergy} \\ \left( \begin{array}{c} \epsilon_d(\mbox{\boldmath $r$}) \\ \epsilon_{\bar{d}}(\mbox{\boldmath $r$}) \end{array} \right) &=& \Omega_q^*(\mbox{\boldmath $r$}) \pm R_j^* \left( V^q_\omega(\mbox{\boldmath $r$}) - \frac{1}{2} V^q_\rho(\mbox{\boldmath $r$}) \right), \label{denergy} \\ \epsilon_{c}(\mbox{\boldmath $r$}) &=& \epsilon_{\bar{c}}(\mbox{\boldmath $r$}) = \Omega_{c}(\mbox{\boldmath $r$}), \label{cenergy} \end{eqnarray} where $\Omega_q^*(\mbox{\boldmath $r$}) {=}\sqrt{x_q^2{+}(R_j^* m_q^*)^2}$, with $m_q^*{=}m_q{-}g^q_\sigma \sigma(\mbox{\boldmath $r$})$ and $\Omega_{c}(\mbox{\boldmath $r$}){=}\sqrt{x_{c}^2{+}(R_j^* m_{c})^2}$. The bag eigenfrequencies, $x_q$ and $x_{c}$, are determined by the usual, linear boundary condition~\cite{Guichon,Saito2}. Note that the lowest eigenenergy for the Dirac equation (Hamiltonian) for the quark, which is positive, can be thought of (for many purposes) as a constituent quark mass. Now, the $D$ and $\bar{D}$ meson masses in the nucleus at position \mbox{\boldmath $r$} (we take $m_D = m_{\bar{D}}$ in vacuum, and then, $m^*_D = m^*_{\bar{D}}$ in nuclear medium), are calculated by: \begin{eqnarray} m_D^*(\mbox{\boldmath $r$}) &=& \frac{\Omega_q^*(\mbox{\boldmath $r$}) + \Omega_c(\mbox{\boldmath $r$}) - z_D}{R_D^*} + {4\over 3}\pi R_D^{* 3} B, \label{md} \\ & &\left. \frac{\partial m_D^*(\mbox{\boldmath $r$})} {\partial R_D}\right|_{R_D = R_D^*} = 0. \label{equil} \end{eqnarray} In Eq.~(\ref{md}), the $z_j$ parametrize the sum of the center-of-mass and gluon fluctuation effects, and are assumed to be independent of density. The parameters are determined in free space to reproduce their physical masses. \begin{figure}[b] \phantom{aa}\vspace{-1.0cm} \psfig{file=comic6.ps,height=12cm,width=9.2cm} \phantom{aa}\vspace{-0.7cm} \caption[]{The $D^-$ and $D^+$ potentials calculated for $^{12}C$ (a) and $^{197}Au$ (b) as a function of the nuclear radius. We also show the downward shift in the threshold for $D^+D^-$ production for $^{12}C$ and $^{197}Au$, in (c).} \label{comic6} \end{figure} In this study we chose the values $m_q{\equiv}m_u{=}m_d{=}5$ MeV and $m_c{=}1300$ MeV for the current quark masses, and $R_N{=}0.8$ fm for the bag radius of the nucleon in free space. Other input parameters and some of the quantities calculated are given in Refs.~\cite{Guichon,Saito2,Tsushima1}. We stress that while the model has a number of parameters, only three of them, $g^q_\sigma$, $g^q_\omega$ and $g^q_\rho$, are adjusted to fit nuclear data -- namely the saturation energy and density of symmetric nuclear matter and the bulk symmetry energy. None of the results for nuclear properties depend strongly on the choice of the other parameters -- for example, the relatively weak dependence of the final results for the properties of finite nuclei, on the chosen values of the current quark mass and bag radius, is shown explicitly in Refs.~\cite{Guichon,Saito2}. Exactly the same coupling constants, $g^q_\sigma$, $g^q_\omega$ and $g^q_\rho$, are used for the light quarks in the mesons as in the nucleon. However, in studies of the kaon system, we found that it was phenomenologically necessary to increase the strength of the vector coupling to the non-strange quarks in the $K^+$ (by a factor of $1.4^2$) in order to reproduce the empirically extracted $K^+$-nucleus interaction~\cite{Tsushima2,Waas}. We assume that the dynamical chiral symmetry breaking for the light quarks in the $D$ and $\bar{D}$ is the same as those for the kaon~\cite{Tsushima1,Tsushima2,Waas}, and will use the stronger vector potential, $\tilde{V}^q_\omega$ $(= 1.4^2 V^q_\omega)$, in this study. \begin{figure}[h] \phantom{aa}\vspace{-0.4cm} \psfig{file=comic7.ps,height=10cm,width=9.cm} \phantom{aa}\vspace{-0.7cm} \caption[]{The total energies of the $D^-$ and $D^+$ mesons at zero momentum calculated for nuclear matter and plotted as function of the baryon density, in units of the saturation density of nuclear matter, $\rho_0{=}0.15$ fm$^{-3}$.} \label{comic7} \end{figure} Through Eqs.~(\ref{diracu}) -- (\ref{equil}) we self-consistently calculate effective masses, $m^*_D(\mbox{\boldmath $r$})$, and mean field potentials, $V^q_{\sigma,\omega,\rho} (\mbox{\boldmath $r$})$, at position $\mbox{\boldmath $r$}$ in the nucleus. The scalar and vector potentials (neglecting the Coulomb force) felt by the $D$ and $\bar{D}$ mesons, which depend only on the distance from the center of the nucleus, $r = |\mbox{\boldmath $r$}|$, are given by: \begin{eqnarray} U^{D^\pm}_s(r) &\equiv& U_s(r) = m^*_D(r) - m_D, \label{spot}\\ U^{D^\pm}_v(r) &=& \mp (\tilde{V}^q_\omega(r) - \frac{1}{2}V^q_\rho(r)), \label{vdpot1} \end{eqnarray} The $\rho$ meson mean field potential, $V^q_\rho(r)$, (and the Coulomb potential) which are small and expected to give a minor effect, will be neglected in the present study. Note that $V^q_\rho$ is negative in a nucleus with a neutron excess. Fig.\ref{comic6} shows the potentials for the $D^-$ and $D^+$-mesons as a function of the nuclear radius calculated for $^{12}C$ and $^{197}Au$. For the following calculations we define the potential as \begin{equation} U^{D^\pm}(r) = U_s(r) + U^{D^\pm}_v(r), \end{equation} where $U_s$ and $U_v$ denote the scalar and vector pieces of the potential, respectively. The in-medium dispersion relation, for the total energy $E_{D^\pm}$ and the momentum $p$ of the $D^\pm$-meson is now given by \begin{equation} E_{D^\pm}(\mbox{\boldmath $r$}) = \sqrt{p^2 + (m_D + U_s(r))^2 } + U^{D^\pm}_v(r), \label{totale} \end{equation} where the bare $D$-meson mass is $m_D{=}1.8693$~GeV. Note that the total $D^-$-meson potential is repulsive, while the $D^+$ potential is attractive, which is analogous to the case for the $K^+$ and $K^-$ mesons, respectively~\cite{Tsushima2}. The amount of downward shift of the $\bar{p}N{\to}D^+D^-$ reaction threshold in nuclei, associated with the in-medium modification of the $D$ and $\bar{D}$ scalar potentials and the vector potentials, is simply $2U_s$, and is shown in Fig.\ref{comic6}c) for $^{12}C$ and $^{197}Au$ as a function of the nuclear radius. The threshold reduction is quite large in the central region of these nuclei and should be detected as an enhanced production of the $D^+D^-$ pairs. Note that a similar situation holds for the $K^+$ and $K^-$ production and, indeed, enhanced $K^-$-meson production in heavy ion collisions, associated with the reduction of the production threshold, has been partially confirmed experimentally~\cite{Kaos,Schroter,Ritman,Barth,Shin,Li,Cassing,Li1,Cassing1}. In Fig.\ref{comic7} the total energies of the $D^+$ and $D^-$ mesons at zero momentum in Eq.~(\ref{totale}), are shown as function of the nuclear matter density in units of normal nuclear matter density ($\rho_0{=}0.15$ fm$^{-3}$). Note that at $\rho_0$ the threshold reduction is around 164~MeV, which should be detectable in $\bar{p}A$ annihilation. \section{The model for D-meson production in ${\rm{\bf\bar{p}}}$A annihilations} The $D\bar{D}$ production in antiproton-nucleus annihilation was calculated using the cascade model~\cite{Sibirtsev1} adopted for $\bar{p}A$ simulations. The detailed description of the initialization procedure as well as the interaction algorithm are given in Ref.~\cite{Sibirtsev1}. The reaction zone was initialized with the use of the momentum dependent $\bar{p}N$ total cross section, given as~\cite{PDG1}: \begin{equation} \sigma_{\bar{p}N}=38.4 + 77.6p^{-0.64}+0.26(\ln{p})^2-1.2\ln{p}, \end{equation} where $p$ denotes the antiproton laboratory momentum and the cross section was taken to be the same for the proton and the neutron target (in good agreement with the data~\cite{PDG1}). The $\bar{p}N{\to}D\bar{D}$ cross section was calculated with quark-gluon string model proposed in Ref.~\cite{Kaidalov3}. In the following we will concentrate on the production of $D^+$ and $D^-$-mesons and thus take into account only two possible reactions, namely $\bar{p}p{\to}D^+D^-$ and $\bar{p}n{\to}D^0D^-$. Note that the relation, \begin{equation} 4\sigma (\bar{p}p\to D^+D^-)= \sigma (\bar{p}n\to D^0D^-) \end{equation} is due to the difference in the number of the quark planar diagrams~\cite{Kaidalov3}. Furthermore, to account for the $D^-$ and $D^+$-meson propagation in nuclear matter one needs to introduce the relevant cross sections for elastic and inelastic $DN$ scattering. Since no data for the $DN$ interaction are available we use a diagrammatic approach illustrated by Fig.\ref{comic14}a,b). Let us compare the $D^-N{\to}D^-N$ and the $K^+N{\to}K^+N$ reactions in terms of the quark lines. Apart from the difference between the $c$ and $s$ quarks, both reactions are very similar and can be understood in terms of rearrangement of the $u$ or $d$ quarks. Thus, in the following calculations we assume that $\sigma_{D^-N{\to}D^-N}{=}\sigma_{K^+N{\to}K^+N}$. \begin{figure}[h] \phantom{aa}\vspace{-0.7cm} \psfig{file=comic14.ps,height=11cm,width=8.4cm} \phantom{aa}\vspace{-1.1cm} \caption[]{Quark diagrams for $K^+n{\to}K^+n$ (a) and $D^-p{\to}D^-p$ (b) elastic scattering and for $K^-n{\to}\Lambda\pi^-$ (c), $D^+p{\to}\Lambda_c^+\pi^+$ (d) inelastic scattering.} \label{comic14} \end{figure} The $K^+N$ cross section was taken from Ref.\cite{Sibirtsev2}, which gives a parametrization of the available experimental data. The total $K^+N$ cross section, averaged over neutron and proton targets, is shown in Fig.\ref{comic3}a) by the dashed line - as a function of the kaon momentum in the laboratory system. Note, that within a wide range of kaon momentum $\sigma_{K^+N}$ is almost constant and approaches a value of ${\simeq}20$~mb. We adopt the value $\sigma_{D^-N}{=}20$~mb, noting that it is entirely due to the elastic scattering channel. Now, Fig.\ref{comic14}c,d) shows both the $K^-N{\to}\Lambda\pi$ and $D^+N \to \Lambda_c\pi$ processes, which are again quite similar in terms of the rearrangement of the $s$ and $c$ quarks, respectively. Thus we assume that $\sigma_{D^+N{\to}\Lambda_c\pi}{\simeq}$ $\sigma_{K^-N{\to}\Lambda\pi}$. \begin{figure}[h] \phantom{aa}\vspace{-0.8cm} \psfig{file=comic3.ps,height=12cm,width=9.cm} \phantom{aa}\vspace{-0.6cm} \caption[]{a) The total $K^-N$ (solid) and $K^+N$ (dashed line) cross sections obtained~\protect\cite{Sibirtsev2} as the best fit to the available experimental data~\protect\cite{PDG} and shown as function of the kaon momentum. The dotted line show the result as explained in the text. b) The $D^-N$ (dashed) and $D^+N$ total cross sections used in the calculations.} \label{comic3} \end{figure} The total $K^-N$ cross section is shown by the solid line in Fig.\ref{comic3}. Again it is averaged over proton and the neutron and taken as a parametrization~\cite{Sibirtsev2} of the experimental data. At low momenta the $K^-N$ cross section shows resonance structures due to the strange baryonic resonances~\cite{PDG}, while at high momenta it approaches a constant value. Apart from the contribution from these intermediate baryonic resonances the inelastic $K^-N{\to}\Lambda\pi$ cross section can be written as \begin{eqnarray} \label{matr} \sigma_{K^-N{\to}\Lambda\pi}=\frac{|M|^2}{16\ \pi \ s} \nonumber \\ \times \left\lbrack \frac{(s-m_\Lambda^2-m_\pi^2)^2-4m_\Lambda^2m_\pi^2} {(s-m_K^2-m_N^2)^2-4m_N^2m_K^2}\right\rbrack^{1/2}, \end{eqnarray} where $s$ is the square of the invariant collision energy and $m_K$, $m_N$, $m_\Lambda$, $m_\pi$ are the masses of kaon, nucleon, $\Lambda$-hyperon and pion, respectively. In Eq.(\ref{matr}) the $|M|$ denotes the matrix element of the $K^-N{\to}\Lambda\pi$ transition, which was taken as a constant. Now the total $K^-N$ cross section is given as a sum of the cross section for the inelastic channel~(\ref{matr}) and for the elastic one, where the latter was taken to be 20.5~mb. The dotted line in Fig.\ref{comic3} shows our result for the total $K^-N$ cross section obtained with $|M|{=}$11.64~GeV$\cdot$fm, which reproduces the trend of the data reasonably well. A similar approach was used to construct the $D^+N$ total cross section. It was assumed that at high momenta the $D^+N$ elastic cross section equals the $D^-N$ cross section, while the $D^+N{\to}\Lambda_c^+\pi$ cross section was calculated from Eq.\ref{matr}, replacing the particle masses as appropriate. The final results are shown in Fig.\ref{comic3} and were adopted for the following calculations. \begin{figure}[b] \phantom{aa}\vspace{-0.7cm} \psfig{file=comic5.ps,height=12cm,width=9.cm} \phantom{aa}\vspace{-0.5cm} \caption[]{The plot of the annihilation zone for $\bar{p}{+}^{12}C$ (a) and $\bar{p}{+}^{197}Au$ (b) reactions at a beam energy of 5 GeV. The solid line indicates the r.m.s. radius of the target nucleus. The arrows show the direction of the antiproton beam.} \label{comic5} \end{figure} We wish to emphasize that the status of the $D$-meson-nucleon interactions is still unknown and is itself one of the important goals of the $\bar{p}A{\to}D\bar{D}X$ studies. Our approach is necessary in order to estimate the expected sensitivity of the experimental measurements to the $DN$ interaction and to study the possibility to evaluate the $D^+N$ and $D^-N$ cross sections. \section{Testing the D-mass in matter} In comparison to low energy antiprotons that annihilate at the periphery of the nucleus because of the large $\bar{p}N$ annihilation probability, antiprotons with energies above 3~GeV should penetrate the nuclear interior. They can therefore probe the nuclear medium at normal baryon density $\rho_0$ and hence yield information about the in-medium properties of the particles. Indeed, as is illustrated by Fig.\ref{comic6}, the $D$-meson potential deviates strongly from zero in the interior of the nuclei considered. Fig.\ref{comic5} shows the reaction zone for the $\bar{p}C$ and $\bar{p}Au$ annihilations at an antiproton beam energy of 5~GeV. The plots are given as a functions of the impact parameter $b$ and the $z$-coordinate, assuming the beam is oriented along the $z$-axis, which is shown by arrows in Fig.\ref{comic5}. The annihilation zone is concentrated in the front hemisphere of the target nuclei. Actually the antiprotons penetrate sufficiently deeply to test densities near that of normal nuclear matter and hence the shift in the $D^+D^-$ production threshold should be manifest. \begin{figure}[h] \psfig{file=comic10.ps,height=12cm,width=9.cm} \phantom{aa}\vspace{-0.7cm} \caption[]{The total cross section for $D^+$ and $D^-$-meson production in $\bar{p}C$ annihilation as a function of the antiproton energy. The results are shown for calculations with free (dashed lines) and in-medium masses (solid lines) for the $D$-mesons. The arrow indicates the reaction threshold on a free nucleon.} \label{comic10} \end{figure} Now we calculate the total $\bar{p}A{\to}D^+D^-X$ production cross section as function of the antiproton beam energy and show the results in Fig.\ref{comic10} for a carbon target and in Fig.\ref{comic1} for gold. The vacuum $\bar{p}N{\to}D^+D^-$ cross section is also shown in Fig.\ref{comic1}. Note that the difference between the $D^+$ and $D^-$-meson production rates is caused by the $D^+$-absorption in nuclear matter. Obviously the production threshold is substantially reduced as compared to the antiproton annihilation on a free nucleon. Apparently, part of this reduction is due to the Fermi motion~\cite{Sibirtsev3,Debowski,Sibirtsev4}, however the calculations with in-medium $D$-meson masses indicate a much stronger threshold reduction comparing to those using the free masses for the final $D$-mesons. Note that, because of their relatively long mean life, the $D$-mesons decay outside the nucleus and their in-medium masses cannot be detected through a shift of the invariant mass of the decay products (unlike the leptonic decay of the vector mesons). Thus it seems that the modification of the $D^+$ and $D^-$-meson masses in nuclear matter can best be detected experimentally as for the shift of the in-medium $K^+$ and $K^-$-meson masses, namely as an enhanced $D$-meson production rate at energies below the threshold for the $\bar{p}N{\to}D^+D^-$ reaction in free space. \begin{figure}[b] \phantom{aa}\vspace{-0.7cm} \psfig{file=comic1.ps,height=12cm,width=9.cm} \phantom{aa}\vspace{-0.8cm} \caption[]{The total cross section for $D^+$ and $D^-$-meson production in $\bar{p}Au$ annihilation as function of the antiproton energy. The results are shown for calculations with free (dashed lines) and in-medium masses (solid lines) for the $D$-mesons. For comparison the vacuum $\bar{p}N{\to}D^+D^-$ cross section is also indicated.} \label{comic1} \end{figure} We should note that experimentally it may be difficult to distinguish whether such an enhancement is due to the modification of the $D^+$ and $D^-$-meson masses in nuclear matter, or due to the Fermi motion, or due to other processes that are not yet included in our study. In principle, the high momentum component of the nuclear spectral function can provide sufficient energy for particle production far below the reaction threshold in free space~\cite{Sibirtsev3}. However, the calculations in Refs.~\cite{Debowski,Sibirtsev4} with realistic spectral functions~\cite{Benhar,Sick,Atti} indicate that such effects are actually negligible, while a more important contribution comes from multistep production mechanism. For instance, the dominant contribution to $K^+$ production in $pA$ collisions comes from the secondary ${\pi}N{\to}YK^+$ process, which prevails over the direct $pN{\to}NYK^+$ reaction mechanism~\cite{Debowski,Badala}. Thus the interpretation of the data depends substantially on the reliable measurement of the production mechanism. It is important, that an additional advantage of the $D^+D^-$ production in $\bar{p}A$ annihilation is the possibility to reconstruct the production mechanism directly. Let us denote as $M_X$ the missing mass of the target nucleon in the reaction $\bar{p}N{\to}D^+D^-$. Obviously, in vacuum $M_X$ is equal to the free nucleon mass and can be reconstructed for antiproton energies above the $D^+D^-$ production threshold on the free nucleon. When analysing $M_X$ in $\bar{p}A$ annihilations one expects the distribution $d\sigma{/}dM_X$ to be centered close to the mass of the bound nucleon - below the free nucleon mass. The shape of the distribution $d\sigma{/}dM_X$ is related to the spectral function of the nucleus~\cite{Benhar,Sick,Atti}. The preceding discussion is based on the hypothesis that the reaction $\bar{p}N{\to}D^+D^-$ is the dominant mechanism for $D^+D^-$ pair production. By measuring both the $D^+$ and $D^-$ mesons one can directly check this hypothesis. \begin{figure}[h] \psfig{file=comic13.ps,height=12cm,width=9.cm} \phantom{aa}\vspace{-0.7cm} \caption[]{The missing mass distribution calculated for $\bar{p}C$ annihilation at 5~GeV. The upper part shows the results obtained without (hatched histogram) and with account of the in-medium potentials (open histogram), but neglecting the $D$-meson interactions in the nucleus. The hatched histogram is normalized to the open histogram. The lower part shows the calculations with $D^+$ and $D^-$ potentials and with $DN$ interactions.} \label{comic13} \end{figure} Let us first neglect the $D$-meson interactions in the nuclear enviroment and analyze the $M_X$ spectrum for $\bar{p}C$ annihilation at 5~GeV. Fig.\ref{comic13}a) shows the missing mass distribution calculated without (hatched histogram) and with inclusion of the $D^+$ and $D^-$-meson potentials. We recall that calculations with free masses provide much smaller $\bar{p}C{\to}D^+D^-X$ production cross sections (see Fig.\ref{comic10}). Thus, for the purpose of the comparison in Fig.\ref{comic13}a) the result obtained without potentials is renormalized to those with in-medium masses. The arrow in Fig.\ref{comic13}a) indicates the density averaged mass of the bound nucleon in the carbon target~\cite{Saito4}. Indeed both histograms are centered around the expected value. However, the calculation with the potentials shows a substantially wider distribution. This effect can be easily understood in terms of the downward shift of the threshold for $D$-meson production in medium. Fig.\ref{comic13}b) shows the $M_X$ distribution calculated with in-medium masses, taking into account both $D^+$ and $D^-$-meson interactions in the nuclear environment. Note that the distribution below $M_X{\simeq}0.75$~GeV results from secondary $DN$ elastic rescattering and its strength is proportional to the $DN$ elastic cross section. A deviation of the actual experimental missing mass distribution from those shown in Fig.\ref{comic13}b) might directly indicate the contribution from $D^+D^-$ reaction mechanisms, other than direct production. In principle, the missing mass, $M_X$, reconstruction appears as a very promising tool for the detection of the in-medium mass modification. Of course, this method requires a detailed knowledge of the nuclear spectral function~\cite{Benhar,Sick,Atti} as well as an accurate calculation of the $M_X$ distribution, which should be compared to the experimental one. \section{Determination of the DN cross section} Obviously both \, $D^+$ and $D^-$ \, mesons are produced in the fragmentation region of the incident antiproton. The $D^+D^-$ pairs gain the total energy available from the $\bar{p}N$ annihilation and because of the high velocity of the antiproton beam they should move forward with large momenta - at least when the produced mesons do not interact with the target. Fig.\ref{comic9} shows the $D^+$ and $D^-$-meson distribution in momentum space, i.e. over the transverse momentum $p_t$, and laboratory, longitudinal momentum, $p_l$, calculated for the $\bar{p}Au$ reaction at an antiproton energy of 7~GeV. The solid lines indicate the $D$-meson emission angle in the laboratory system. The $D$-meson distribution in momentum space shows two branches. The branch at large $p_l$ and small $p_t$ originates from the primary production of the $D^+D^-$ pairs in the antiproton annihilation at the target nuclei. The width of this branch reflects the spectral function of the nuclei, i.e. the internal momentum and energy distribution of the nucleons~\cite{Benhar,Sick,Atti}. \begin{figure}[h] \phantom{aa}\vspace{-0.8cm} \psfig{file=comic9.ps,height=12cm,width=8.7cm} \phantom{aa}\vspace{-0.8cm} \caption[]{The distribution over the transverse $p_t$ and laboratory longitudial momentum $p_l$ for the $D^-$ (upper part) and $D^+$-meson (lower part) produced in $\bar{p}Au$ annigilations at 7~GeV. Lines indicates the detection angle in the laboratory system.} \label{comic9} \end{figure} The second branch in Fig.\ref{comic9} is located at small $p_l$ and originates from the elastic and inelastic interactions of the $D$-meson in nuclear medium. Note that the $D^+$ mesons are produced in the annihilation with sufficiently large momenta that they are not strongly absorbed (see Fig.\ref{comic3}) but can be scattered elastically similar to the $D^-$-mesons. Obviously, to be absorbed the $D^+$-mesons should first be slowed down in the nuclear matter. Thus the experimental study of the charm exchange reaction, $D^+N{\to}\Lambda_c^+\pi$, seems to be more informative with the heavy nuclear targets, where multiple scattering is more probable. The momentum spectra for $D^-$ and $D^+$-mesons produced in $\bar{p}C$ annihilation at 5~GeV are shown in Fig.\ref{comic8}. The hatched histograms are the primary spectra from $\bar{p}$-annihilation on the target nucleon, while the solid histograms show the final $D$-meson spectra. The difference between the primary and final spectra arise primary from elastic rescattering. For such a light target as carbon, the $D^+$-absorption is almost negligible and therefore the difference between the $D^-$ and $D^+$ momentum spectra is only the absolute normalization. Note that the $D^+$ can be produced in $\bar{p}$ annihilation at the target proton, while $D^-$ - can be produced on either a neutron or proton. \begin{figure}[h] \phantom{aa}\vspace{-0.7cm} \psfig{file=comic8.ps,height=12.4cm,width=8.7cm} \phantom{aa}\vspace{-0.6cm} \caption[]{The momentum spectra of $D^-$ and $D^+$ mesons in the laboratory system and from the $\bar{p}C$ annihilations at 5 GeV. Hatched histograms show the primary spectra from the antiproton annihilation at the bound nucleon. Solid histograms are the final spectra.} \label{comic8} \end{figure} A rather different situation applies for the antiproton annihilation on heavy targets. Fig.\ref{comic2} shows the $D$-meson spectra from $\bar{p}Au$ annihilations at 5~GeV. Again the hatched histograms are the primary spectra from the annihilation, while the solid histograms show the final spectra. The $D^-$-meson spectrum is enhanced at low momenta which indicates strong $D^-$ deceleration in the gold. At the same time the total $D^-$ yield does not change in comparison with the primary production. The $D^+$-spectrum is substantially different from the primary (shadowed histogram) and around 40\% of the initial $D^+$-mesons, produced in $\bar{p}Au$ annihilation at 5 GeV, are absorbed. Indeed the difference between the $D^+$ and $D^-$ spectra comes from the $D$-nucleon absorption. \begin{figure}[b] \phantom{aa}\vspace{-0.5cm} \psfig{file=comic2.ps,height=12cm,width=8.7cm} \phantom{aa}\vspace{-0.8cm} \caption[]{The momentum spectra of $D^-$ and $D^+$ mesons in the laboratory system and from the $\bar{p}Au$ annihilations at 5 GeV. Hatched histograms show the primary spectra from the antiproton annihilation at the bound nucleon. Solid histograms are the final spectra.} \label{comic2} \end{figure} \section{Conclusion} We have studied $D$-meson production in antiproton-nucleus annihilation. It was found that $\bar{p}A$ annihilation at energies below the $\bar{p}N{\to}D^+D^-$ reaction threshold in free space offer reasonable conditions for the detection of the changes in $D$-meson properties in-medium at normal nuclear matter density. In-medium modification of the $D$-meson mass can be observed as an enhanced $D^+D^-$ production at antiproton energies below ${\simeq}5.5$~GeV. The advantage of the $\bar{p}A{\to}D^+D^-X$ reaction is the possibility to reconstruct directly the primary production mechanism and hence to avoid a mistaken interpretation of such an enhancement as due to the contribution from multistep production processes. In part this reconstruction allows one to restrict the data analysis in terms of the effect due to the high momentum component of the nuclear spectral function. The study of the in-medium modification of the $D$-meson mass seems very promising, even with a target as light as carbon, where the total $D^+D^-$ mass reduction is sizeable and the nuclear spectral function is under control~\cite{Benhar,Sick,Atti}. We found that the \, $\bar{p}A$ \, annihilations also provide favourable conditions for studying the $D$-meson interaction in the nucleus. The difference in the $D^+$ and $D^-$-meson momentum spectra from antiproton annihilation on heavy nuclei provides a very clean signature for the charm exchange $D^+N{\to}\Lambda_c^+\pi$ reaction and can be used for the determination of the $DN$ rescattering and absorption cross sections. \acknowledgement{ A.S would like to acknowledge the warm hospitality and partial support of the CSSM. This work was supported by the Australian Research Council and the Forschungzentrum J\"{u}lich.}
\section{Introduction} \label{section:introduction} The experimental search \cite{Higgsexp} for the Higgs \cite{Higgs} boson is mainly concerned nowadays by that of its strictly standard (unique) avatar \cite{GlashowSalamWeinberg} or of its multiple supersymmetric \cite{SUSY} reincarnations. The present and next generations of accelerators should be able to give us precious information about this cornerstone of particle physics. If a positive outcome would undoubtedly be one of the greatest achievements of the century and provide a fair reward for the simplest and most elegant ideas, a negative one should not be considered as a failure and be interpreted as the absence of the Higgs but rather as an encouragement to refine our ideas, to look for alternatives, and in particular to investigate whether one could have missed a Higgs particle with couplings smaller than predicted in the basic models. This is the goal of this work, which gives an example of a nearly standard Higgs boson, still considered to be unique, which could have escaped detection. I outpass here the present negative feelings attached to technicolour models \cite{technicolour} (mainly due to the problem of flavour changing neutral currents) and revive a composite Higgs boson; however, if it does transform like a quark-antiquark pair, it nevertheless appears, unlike the quarks, as a fundamental field in the Lagrangian, together with the observed pseudoscalar (and scalar) mesons, along the lines of \cite{Machet1}. It is a neat way of including both chiral and electroweak properties of asymptotic states. As the proposed model respects in particular the standard electroweak transformations of quarks, the customary problems associated with flavour changing neutral currents do not arise (we shall see that a new type of them occurs but only in the Higgs sector, and beyond our present detection capabilities). The only concern being here the electroweak physics of scalar (Higgs) and pseudoscalar $J=0$ states, Quantum Chromodynamics (QCD) \cite{QCD} and strong interactions are deliberately left aside, except in that they determine the asymptotic mesons and their normalization; interactions among particles in the final state are not considered. The quark picture is often invoked, but only to make the reader more comfortable and to provide a link with customary considerations: the quarks themselves, which are not asymptotic fields, do not appear in the Lagrangian. The breakdown of the electroweak symmetry, tantamount to a condensation of quark-antiquark pairs, is treated on a purely phenomenological ground (much alike in the standard electroweak model) without reference to a deeper underlying mechanism. The Higgs boson being searched for through its decays, determining its coupling to {\it observed} particles is essential. Yukawa couplings to quarks are now absent, but the Higgs couples through a ``mexican hat'' potential to the three goldstones of the broken electroweak symmetry, which are, as is shown by the study of their leptonic decays, linear combinations of pseudoscalar mesons. There exist $N^2/4$ ($N/2$ is the number of generations) real electroweak quadruplets of scalar and pseudoscalar (composite) fields isomorphic to the complex isoscalar doublet of the Glashow-Salam-Weinberg model \cite{GlashowSalamWeinberg}. Accordingly, the flavour orientation of the Higgs boson is determined by a rotation in this $N^2/4$ dimensional space. It determines, too, the flavour content of the three (pseudoscalar) Goldstone bosons which are the three partners of the Higgs in the same quadruplet, and the hierarchy between the leptonic decay constants $f$ of the various pseudoscalar mesons. This makes possible the determination of the couplings of the Higgs to pseudoscalar mesons as a function of the $f$'s, of Cabibbo-Kobayashi-Maskawa (CKM) \cite{CKM} mixing angles and of the mass of the Higgs itself. Its couplings to gauge fields keep as in the standard model. I also show how, in an interpretation in terms of quarks, the orientation of the Higgs boson determines the hierarchy of the different quark-antiquark vacuum expectation values. The paper is organized as follows:\l - in section \ref{section:electroweak}, I recall the theoretical basis of the model;\l - in section \ref{section:EWbreaking}, I study the orientation of the Higgs boson in flavour space and its consequences;\l - in section \ref{section:decays}, I specialize to the three decays $Z \rightarrow e^+e^- B^+B^-$, $Z \rightarrow e^+e^- D_s^+ D_s^-$ and $Z \rightarrow e^+e^- K^\pm \pi^\mp$;\l - finally, appendix \ref{app:Dmatrix} gives technical details concerning the basis of electroweak and flavour eigenstates, and appendix \ref{app:GMOR} provides a link between the normalization of the asymptotic mesons and the Gell-Mann-Oakes-Renner relation \cite{GellMannOakesRenner} in QCD. \section{Electroweak interactions of quark-antiquark composite fields} \label{section:electroweak} \subsection{Electroweak and ``flavour'' eigenstates} \label{subsec:eigenstates} Though we only deal here with electroweak physics, the paper rests on the fact that the ``asymptotic'' states for $J=0$ mesons can be interpreted as ``flavour'' eigenstates, determined by strong interactions \footnote{Strong interactions are considered to be independent of ``flavour'', and so both have common eigenstates.}. This can be put in correspondence with the existence of two different mass scales, hence two different characteristic times:\l - the electroweak mass scale $M_W \approx 80$ GeV, with an associated time scale $\tau_{EW} = 1/M_W$;\l - the mass scale associated with strong interactions, with an order of magnitude of the masses of the mesons and resonances exchanged in nuclear interactions, that is a few hundred MeV; the associated time scale $\tau_S$ is much larger than $\tau_{EW}$; thus, if an electroweak eigenstate is produced (by electroweak interactions) at time $t$ it can only be detected as such between $t$ and $t + \tau_{EW}$; after this interval, and before it eventually decays into final states which can be non-hadronic, one only detects its flavour components. The meaning of ``asymptotic'', which has to be adapted to the type of problem that is being analyzed, is consequently here ``for time scales larger than $\tau_{EW}$''. While the customary procedure is to try to incorporate strong interactions between asymptotic electroweak eigenstates to build up observed particles, which of course, like when introducing gluonic corrections in QCD, faces non-perturbative problems, I will rather here consider as perturbative and small the electroweak interactions between flavour asymptotic states which are determined by strong interactions. The only additional (non-perturbative) effect of strong interactions that I will introduce is the normalization of asymptotic mesons which is determined from their leptonic decays and is shown to be in agreement with the ``Partially Conserved Current Hypothesis'' (PCAC) and with the Gell-Mann-Oakes-Renner relation in QCD (see subsections \ref{subsubsec:flavour}, \ref{subsubsec:norm}, appendix \ref{app:GMOR}) \footnote{ The question can be raised whether the Higgs can appear as an asymptotic state or is projected at ``large'' times on flavour (scalar) eigenstates; its eventual detectability as an asymptotic particle can depend on it. \label{foot:asympt}} . \subsection{Theoretical framework} \label{subsec:theory} The general framework has been set in \cite{Machet1}. For the sake of understandability and for this paper to be self-contained, I briefly recall here the main useful steps, in a somewhat less formal approach more usable for phenomenological purposes. Quarks are considered to be mathematical objects \cite{GellMann} which are determined by their quantum numbers and by their transformations by the different groups of symmetry that act upon them; we are mainly concerned here with the chiral group $U(N)_L \times U(N)_R$ where $N$ is the number of ``flavours'', and with the electroweak group $SU(2)_L \times U(1)$: they form an $N$-vector $\Psi$ \begin{equation} \Psi = \left( \begin{array}{c} u\\ c\\ \vdots \\d\\ s\\ \vdots \end{array} \right) \label{eq:Psi}\end{equation} in the fundamental representation of the diagonal subgroup of the chiral group, and their electroweak transformations, to which we stick to, are the usual ones of the Glashow-Salam-Weinberg model \cite{GlashowSalamWeinberg}. Any $SU(2)_L \times U(1)$ group can be considered, for $N$ even, as a subgroup of $U(N)_L \times U(N)_R$; that it be the electroweak group, {\it i.e.} that it act on quarks in the standard way determines its embedding in the chiral group. It is easy to check that the $SU(2)$ group the three generators of which are the three $N \times N$ matrices (they are written in terms of four $N/2 \times N/2$ sub-blocs with $\mathbb I$ the unit matrix and $\mathbb K$ the CKM mixing matrix) \begin{equation} {\mathbb T}^3_L = {1\over 2}\left(\begin{array}{rrr} {\mathbb I} & \vline & 0\\ \hline 0 & \vline & -{\mathbb I} \end{array}\right),\ {\mathbb T}^+_L = \left(\begin{array}{ccc} 0 & \vline & {\mathbb K}\\ \hline 0 & \vline & 0 \end{array}\right),\ {\mathbb T}^-_L = \left(\begin{array}{ccc} 0 & \vline & 0\\ \hline {\mathbb K}^\dagger & \vline & 0 \end{array}\right), \label{eq:SU2L} \end{equation} acting trivially on the left-handed projection $\Psi_L = [(1-\gamma_5)/2] \Psi$ of $\Psi$ can be identified with the standard electroweak $SU(2)_L$; the $U(1)$ associated to the weak hypercharge $\mathbb Y$ is determined through the Gell-Mann-Nishijima relation \cite{GellMannNijishima} \begin{equation} ({\mathbb Y}_L,{\mathbb Y}_R) = ({\mathbb Q}_L,{\mathbb Q}_R) - ({\mathbb T}^3_L,0), \label{eq:GMN}\end{equation} and from the trivial form for the (diagonal) charge operator $\mathbb Q$ \begin{equation} {\mathbb Q}_L ={\mathbb Q}_R ={\mathbb Q}=\left(\begin{array}{ccc} 2/3 & \vline & 0\cr \hline 0 & \vline & -1/3 \end{array}\right). \label{eq:Q} \end{equation} The $2N^2$ composite of the form $\bar q q$ or $\bar q\gamma_5 q$ can be cast into $N^2/2$ quadruplets which are stable by the electroweak group; their flavour structure is materialized by $N\times N$ matrices $\mathbb M$ and the quadruplets can generically be written \vbox{ \begin{eqnarray} & &\Phi(\mathbb D)= ({\mathbb M}\,^0, {\mathbb M}^3, {\mathbb M}^+, {\mathbb M}^-)(\mathbb D)\cr & &\ \cr & & =\left[ \frac{1}{\sqrt{2}}\left(\begin{array}{ccc} {\mathbb D} & \vline & 0\\ \hline 0 & \vline & {\mathbb K}^\dagger\,{\mathbb D}\,{\mathbb K} \end{array}\right), \frac{i}{\sqrt{2}} \left(\begin{array}{ccc} {\mathbb D} & \vline & 0\\ \hline 0 & \vline & -{\mathbb K}^\dagger\,{\mathbb D}\,{\mathbb K} \end{array}\right), i\left(\begin{array}{ccc} 0 & \vline & {\mathbb D}\,{\mathbb K}\\ \hline 0 & \vline & 0 \end{array}\right), i\left(\begin{array}{ccc} 0 & \vline & 0\\ \hline {\mathbb K}^\dagger\,{\mathbb D} & \vline & 0 \end{array}\right) \right],\cr & & \label{eq:reps} \end{eqnarray} } where $\mathbb D$ is a real $N/2 \times N/2$ matrix (see subsection \ref{subsubsec:invariants} and Appendix \ref{app:Dmatrix}). One may furthermore consider quadruplets the entries of which have a definite parity (the $\mathbb S$'s below stand for scalars and the $\mathbb P$'s for pseudoscalars) \begin{equation} \varphi = ({\mathbb S}^0, \vec {\mathbb P}), \label{eq:SP} \end{equation} and \begin{equation} \chi = ({\mathbb P}\,^0, \vec {\mathbb S}). \label{eq:PS} \end{equation} The $\varphi$'s and the $\chi$'s transform alike by the gauge group, according to ($i$ and $j$ are $SU(2)$ indices) \vbox{ \begin{eqnarray} {\mathbb T}^i_L\,.\,{\mathbb M}^j &=& -\frac{i}{2}\left( \epsilon_{ijk} {\mathbb M}^k + \delta_{ij} {\mathbb M}^0 \right),\cr {\mathbb T}^i_L\,.\,{\mathbb M}^0 &=& \frac{i}{2}\; {\mathbb M}^i. \label{eq:actioneven} \end{eqnarray} } The link between the matrices $\mathbb M$ and diquark operators is straightforwardly established by sandwiching the latter between $\overline\Psi$ and $\Psi$ and inserting a $\gamma_5$ when needed by parity. The link between diquark operators, of dimension $[mass]^3$, and the scalar (pseudoscalar) fields (electroweak eigenstates) of dimension $[mass]$ occurring in the Lagrangian is achieved by introducing an appropriate normalization as follows. Let $H = h + \langle H\rangle$ be the Higgs boson such that \begin{equation} \langle H \rangle = \frac{v}{\sqrt{2}} \label{eq:Hvev} \end{equation} breaks the electroweak $SU(2)_L \times U(1)$ into its electromagnetic $U(1)_{em}$ subgroup. Its flavour content is represented by an $N\times N$ matrix $\mathbb H$ and the associated diquark operator is $\overline\Psi {\mathbb H} \Psi$. Eq.~(\ref{eq:Hvev}) is trivially satisfied by \begin{equation} H = \frac{\langle H \rangle}{\langle \overline\Psi{\mathbb H}\Psi \rangle} \overline\Psi {\mathbb H} \Psi \label{eq:Hnorm} \end{equation} {from} which one can choose for all fields $\phi$ with dimension $[mass]$ associated with the $N\times N$ matrix $\mathbb M$ the normalization \begin{equation} \phi_{\mathbb M} = (i)\,\frac{\langle H \rangle}{\langle \overline\Psi{\mathbb H}\Psi \rangle} \overline\Psi (\gamma_5){\mathbb M} \Psi. \label{eq:phinorm} \end{equation} \subsubsection{Flavour eigenstates} \label{subsubsec:flavour} Because of the CKM rotation, the electroweak eigenstates $\varphi$ and $\chi$ defined in (\ref{eq:SP},\ref{eq:PS}) are not flavour eigenstates but linear combinations of them. The flavour or ``strong'' eigenstates are the ones associated with $\mathbb M$ matrices which have only one nonvanishing entry equal to $1$. Let ${\mathbb M}_{ab}$ such a matrix with one single non-vanishing entry at the crossing of the $a$-th line and $b$-th column. The associated flavour eigenstate, that we call $\Pi_{ab}$ is, according to (\ref{eq:phinorm}), and in the case of a pseudoscalar \begin{equation} \Pi_{ab}({\mathbb M}_{ab}) = i\,\frac{\langle H \rangle}{\langle \overline\Psi{\mathbb H}\Psi \rangle} \overline\Psi \gamma_5{\mathbb M}_{ab} \Psi. \label{eq:flavoureigenstates} \end{equation} The $\Pi_{ab}$'s are related to asymptotic states, {\it i.e.} observed mesons ${\cal P}_{ab}$, by a scaling factor $\textswab b$ (see subsection \ref{subsubsec:norm}). \subsubsection{Quadratic invariants and electroweak mass scales} \label{subsubsec:invariants} To every quadruplet $({\mathbb M}^0, \vec{\mathbb M})$ is associated a quadratic invariant: \begin{equation} {\cal I} = ({\mathbb M}^0, \vec {\mathbb M})\otimes ({\mathbb M}^0, \vec {\mathbb M}) = {\mathbb {\mathbb M}}\,^0 \otimes {\mathbb {\mathbb M}}\,^0 + \vec {\mathbb M} \otimes \vec {\mathbb M}; \label{eq:invar} \end{equation} the ``$\otimes$'' product is a tensor product (not the usual multiplication of matrices) and means the product of fields as functions of space-time; $\vec {\mathbb M} \otimes \vec {\mathbb M}$ stands for $\sum_{i=1,2,3} {\mathbb M}\,^i \otimes {\mathbb M}\,^i$. For the relevant cases $N=2,4,6$, there exists a set of $\mathbb D$ real matrices (see appendix \ref{app:Dmatrix}) such that the algebraic sum of invariants specified below, extended over all representations defined by (\ref{eq:SP},\ref{eq:PS},\ref{eq:reps}) \begin{equation} {1\over 2} \left((\sum_{symmetric\ {\mathbb D}'s} - \sum_{antisym\ {\mathbb D}'s}) \left( ({\mathbb S}^0, \vec {\mathbb P})({\mathbb D}) \otimes ({\mathbb S}^0, \vec {\mathbb P})({\mathbb D}) - ({\mathbb P}^0, \vec {\mathbb S})({\mathbb D}) \otimes ({\mathbb P}^0, \vec {\mathbb S})({\mathbb D}) \right)\right) \label{eq:diaginvar} \end{equation} is diagonal both in the electroweak basis and in the basis of flavour eigenstates. With the coefficient $(1/2)$ chosen in (\ref{eq:diaginvar}) in the electroweak basis, the normalization in the basis of flavour eigenstates is $(+1)$, with all signs positive. {From} the property stated above, to each quadruplet can be associated an arbitrary electroweak mass scale and, for such a choice of $\mathbb D$ matrices, the degeneracies of electroweak and flavour eigenstates coincide. \subsubsection{The electroweak Lagrangian} \label{subsubsec:lagrangian} The scalar (pseudoscalar) electroweak fields that build up the Lagrangian are taken to be the ones associated with the set of matrices $\mathbb D$ diagonalizing the invariant (\ref{eq:diaginvar}) in both the electroweak and the flavour basis, and the combination used for the kinetic terms is the one of (\ref{eq:diaginvar}). The kinetic terms for the leptons and the gauge fields are the standard ones. No Yukawa coupling to quarks is present since they are not fields of the Lagrangian, and masses are given in a gauge invariant way to the mesons themselves \footnote{The hadronic sector being anomaly-free, a purely vectorial theory in the leptonic sector is then favoured; the approach proposed in \cite{BellonMachet} is an example of that, where leptons are given masses without introducing Yukawa couplings and where the standard model appears as an effective theory when one neutrino helicity decouples by becoming infinitely massive.} . A ``mexican hat'' potential is phenomenologically introduced to trigger the spontaneous symmetry breaking of the electroweak symmetry. \subsubsection{Normalizing the fields} \label{subsubsec:norm} By convention, we take all electroweak eigenstates $\phi$ occurring in the Lagrangian and defined by (\ref{eq:phinorm}) to be normalized to ``$1$'': \begin{equation} \int \frac{d^4p}{(2\pi)^3}\vert\phi(p)\rangle\langle\phi(p) \vert \theta(p_0) \delta(p^2 - m_\phi^2) = 1, \label{eq:norm1} \end{equation} \begin{equation} \langle\phi(p')\vert\phi(p)\rangle = 2p_0 (2\pi)^3 \delta^3(p-p'), \label{eq:norm2} \end{equation} and, for fields corresponding to two different $\mathbb D$ matrices \begin{equation} \langle\phi({\mathbb D}_\alpha)\vert \phi({\mathbb D}_\beta)\rangle = 0\ , \ \alpha\not= \beta. \label{eq:norm3} \end{equation} The phase-space measure for $\phi$ is \begin{equation} d\mu_{\phi(p)} = \frac{d^4p}{(2\pi)^3} \theta(p_0) \delta(p^2 - m_\phi^2). \label{eq:measure1} \end{equation} The flavour eigenstates $\Pi_{ab}$ defined in (\ref{eq:flavoureigenstates}) are then normalized to $(1/2)$ \begin{equation} \langle\Pi(p')\vert\Pi(p)\rangle = (\frac{1}{2})2p_0 (2\pi)^3 \delta^3(p-p'); \label{eq:normPi} \end{equation} the difference of normalization between the $\phi$'s and the $\Pi$'s comes from the identity \begin{equation} \sum_\alpha {\mathbb D}_\alpha \otimes {\mathbb D}_\alpha = 2 \sum_{ab} {\mathbb D}_{ab} \otimes {\mathbb D}_{ab} \end{equation} where the ${\mathbb D}_\alpha$'s are the ones exhibited in appendix \ref{app:Dmatrix} and the ${\mathbb D}_{ab}$'s are $N/2 \times N/2$ matrices the only non-vanishing entry of which is ``$1$'' at the crossing of the $a$-th line and $b$-th column; this property also reflects into the mismatch in the normalizations when expressing the quadratic invariant (\ref{eq:diaginvar}) in terms of electroweak or flavour eigenstates, as stated in subsection \ref{subsubsec:invariants}. The last point concerns the normalization of the asymptotic (observed) mesons ${\cal P}_{ab}$ (${\cal P}^+_{ud} =\pi^+, {\cal P}^-_{su} = K^-\ldots $) which have the same flavour structure as the $\Pi_{ab}$'s: one introduces the scaling factor $\textswab b$ such that \begin{equation} \Pi = {\textswab b} {\cal P}. \label{eq:bdef} \end{equation} For the ${\cal P}$'s one has \begin{equation} \langle {\cal P}(p')\vert {\cal P}(p)\rangle = \frac{1}{2\,{\textswab{b}^2}}2p_0 (2\pi)^3 \delta^3(p-p') \label{eq:norm4} \end{equation} and \begin{equation} d\mu_{{\cal P}(p)} = 2\,{\textswab{b}^2}\frac{d^4p}{(2\pi)^3} \theta(p_0) \delta(p^2 - m_{\cal P}^2). \label{eq:measure2} \end{equation} The scaling factor $\textswab{b}$ is determined from the leptonic decays of the pseudoscalar mesons and is \footnote{In previous works \cite{Machet2} the asymptotic mesons were normalized to ``$1$'', which led to a scaling factor $a = 1/\textswab{b}$. Though the physical results turn out, as expected to be the same, the procedure used here is more systematic and makes the links more conspicuous with the traditional picture of QCD.} \begin{equation} \textswab{b} = \frac{\langle H\rangle}{2f_0} \label{eq:b} \end{equation} where $f_0$ is the generic leptonic decay constant. As a consequence, one has for example (for $N=4$) \begin{equation} P^+({\mathbb D}_1) =\textswab{b} \left(\cos\theta_c(\pi^+ + D_s^+) + \sin\theta_c(K^+ - D^+)\right), \label{eq:projection} \end{equation} where $\theta_c$ is the Cabibbo angle. I show in Appendix \ref{app:GMOR} how this scaling is consistent, through PCAC, with the Gell-Mann-Oakes-Renner relation \cite{GellMannOakesRenner} in QCD. The scaling factor $\textswab b$, that we introduce in a non-perturbative way, and which ensures the consistency with other approaches and with experimental observations, can be put in parallel with the $\sqrt{Z}$'s normalizing asymptotic fields in the $S$-matrix theory; our hypothesis lies here in that, once admitted that strong interactions determine asymptotic states, their only other effect can be phenomenologically parameterized by $\textswab b$ \footnote{We in particular do not consider here the phase-shifts introduced by strong interactions among final states.}. \section{Electroweak symmetry breaking and the Higgs boson} \label{section:EWbreaking} \subsection{The flavour orientation of the Higgs boson} \label{subsec:Higgs} The real $(H,\vec G)$ quadruplet (complex doublet) of the standard model, where $H$ is the (scalar) Higgs boson and $\vec G = (G^+, G^3, G^-)$ are the three goldstones of the broken electroweak symmetry is isomorphic to any of the $N^2/4$ quadruplets $\varphi$ of eq.~(\ref{eq:SP}) \footnote{ Considering that the Higgs boson is unique prevents the occurrence of a hierarchy problem \cite{GildenerWeinberg}} . We thus face an arbitrariness in the flavour content of the Higgs boson, which also determines the composition of the three goldstones in terms of flavour eigenstates since they are its three pseudoscalar partners in the same quadruplet. Identifying $H$ with ${\mathbb S}^0({\mathbb D}_1)$ as I did in previous works \cite{Machet2} is tantamount to taking the same value for all diagonal vacuum expectation values $\langle \bar q_a q_a\rangle$ ($a$ is a flavour index). I loosen here this hypothesis and introduce a real orthogonal rotation matrix $\cal{R}$ acting in the $N^2/4$ dimensional space of the $\varphi$'s. For the sake of simplicity, I will perform the analysis in the case of two generations ($N=4$). We can then restrict furthermore $\cal{R}$ to be a $3\times 3$ rotation matrix by postulating that $\langle {\mathbb S}^0({\mathbb D}_4)\rangle =0$; this is equivalent to saying that $\langle \bar q_a q_b - \bar q_b q_a\rangle =0$, which is true if $CP$ is an unbroken symmetry. So, the $(H,\vec G)$ multiplet is now considered to be a linear combination of $\varphi_1=\varphi({\mathbb D}_1)$, $\varphi_2=\varphi({\mathbb D}_2)$ and $\varphi_3=\varphi({\mathbb D}_3)$ (see eq.~(\ref{eq:SP})), and the $4\times 4$ flavour matrix associated to $G^+$ reads \begin{equation} {\mathbb G}^+ = i\;\left(\begin{array}{ccc} 0 & \vline & {\begin{array}{cc}G^+_{ud} & G^+_{us}\cr G^+_{cd} & G^+_{cs} \end{array}} \\ \hline 0 & \vline & 0 \end{array}\right). \label{eq:G+} \end{equation} Let the orthogonal matrix \footnote{ The orthogonality of $\cal R$ preserves the property stated in subsection \ref{subsubsec:invariants} that the quadratic combination (\ref{eq:diaginvar}) is diagonal for both electroweak and flavour eigenstates; suppose then that there is a single electroweak mass scale $M$ except the vanishing one generated by the breaking of the electroweak symmetry; since, would there be no goldstone, all flavour eigenstates would have the same mass $M$, the ${\text(mass)}^2$ of the flavour components of the charged goldstone $G^+$ are \begin{eqnarray} M_{\pi^\pm}^2 &=& \frac{M^2}{2}(2- (G^+_{ud})^2),\cr M_{K^\pm}^2 &=& \frac{M^2}{2}(2- (G^+_{us})^2),\cr M_{D^\pm}^2 &=& \frac{M^2}{2}(2- (G^+_{cd})^2),\cr M_{D_s^\pm}^2 &=& \frac{M^2}{2}(2- (G^+_{cs})^2). \label{eq:masses} \end{eqnarray} Our statement that strong interactions determine asymptotic states for mesons can thus participate to creating a mass hierarchy for the latter as a consequence of the breaking of the electroweak symmetry. Of course, electroweak interactions among asymptotic states can occur, in particular through the non-diagonal mass terms which do not cancel any longer after the symmetry is broken. } $\cal R$ such that \begin{equation} \left( \begin{array}{c} \tilde\varphi_1 \\ \tilde\varphi_2 \\ \tilde\varphi_3 \end{array} \right) \equiv \left( \begin{array}{c} (\tilde{\mathbb S}_1, \vec{\tilde{\mathbb P}_1}) \\ (\tilde{\mathbb S}_2, \vec{\tilde{\mathbb P}_2}) \\ (\tilde{\mathbb S}_3, \vec{\tilde{\mathbb P}_3}) \end{array} \right) = {\cal R} \left( \begin{array}{c} \varphi({\mathbb D}_1) \\ \varphi({\mathbb D}_2) \\ \varphi({\mathbb D}_3) \end{array} \right) \label{eq:rotation} \end{equation} depend on three mixing angles $\theta_1, \theta_2, \theta_3$, the sines and cosines of which will be noted $s_1,s_2,s_3$ and $c_1,c_2,c_3$ \begin{equation} {\cal R} = \left( \begin{array}{ccc} c_1 & -s_1c_3 & -s_1s_3 \\ s_1c_2 & c_1c_2c_3 - s_2s_3 & c_1c_2s_3 + s_2c_3 \\ s_1s_2 & c_1s_2c_3 + c_2s_3 & c_1s_2s_3 - c_2c_3 \end{array} \right). \label{eq:R} \end{equation} We choose by convention \footnote{This makes in particular our results independent of $\theta_2$, in relation with the fact that one relative phase between the quadruplets has no physical significance.} \begin{equation} (H, \vec G) = \tilde\varphi_1 = (\tilde{\mathbb S}_1, \vec{\tilde{\mathbb P}_1}), \label{eq:defHG} \end{equation} leading to \begin{eqnarray} G^+_{ud} &=& c_\theta(c_1 - s_1s_3) + s_\theta s_1s_3, \\ G^+_{us} &=& s_\theta(c_1 - s_1s_3) - c_\theta s_1s_3, \\ G^+_{cd} &=& -s_\theta(c_1 + s_1s_3) - c_\theta s_1s_3, \\ G^+_{cs} &=& c_\theta(c_1 + s_1s_3) - s_\theta s_1s_3, \label{eq:Gij} \end{eqnarray} where $s_\theta$ and $c_\theta$ stand for the sine and cosine of the Cabibbo angle. \subsection{Leptonic decays and the hierarchy of decay constants} \label{subsec:leptonic} I show here how the orientation of the Higgs boson in flavour space determines the hierarchy of the pseudoscalar leptonic decay constants. The leptonic decays are described by the diagram of Fig.~1; because the non-diagonal $W-\text{meson}$ coupling only occurs for the three goldstones $\vec G$, their flavour components are the only pseudoscalar mesons which can decay leptonically. These decays determine the scaling factor $\textswab b$ (\ref{eq:bdef},\ref{eq:b}) of the flavour eigenstates. \vbox{ \vskip .5cm plus 3mm minus 2mm \begin{center} \epsfig{file=Fig1.ps} \vskip .5cm plus 3mm minus 2mm {\em Fig.~1: The leptonic decay of a pseudoscalar meson.} \end{center} \vskip .5cm plus 3mm minus 2mm } This section gives us the opportunity to explicitly compute a decay amplitude involving asymptotic fields (pseudoscalar mesons) with a normalization different from ``$1$'', while the leptons and gauge fields are normalized to ``$1$'' \footnote{While the $\textswab b$ factor can be reabsorbed by a simple rescaling of the mesonic fields when the Lagrangian involves only mesons, like for example in the non-linear $\sigma$-model, this is more problematic when several types of fields are involved which interact between each other and the different kinetic terms of which come out with different normalizations.} . The calculation proceeds by using a $W$ propagator in the unitary gauge, and we do it here in analogy with the standard ``PCAC'' computation. We have to evaluate \begin{equation} _{out}\langle e^+ \nu_e \vert \pi^+\rangle _{in} = \langle e^+ \nu_e \vert \frac{G_F}{\sqrt{2}}\ L_\mu H^\mu\vert \pi^+\rangle, \label{eq:amplitude1} \end{equation} where $L_\mu$ and $H_\mu$ are respectively the weak leptonic and hadronic currents. $H_\mu$ is deduced from the part of the Lagrangian corresponding to the $(H,\vec G)$ quadruplet \begin{equation} {\cal L} \ni -\frac{gv}{2\sqrt{2}}\left( \partial^\mu G^+W_\mu^- + \partial^\mu G^-W_\mu^+ \right)+ \ldots \label{eq:Lgoldstone} \end{equation} such that, by inserting the vacuum $\vert 0\rangle\langle 0 \vert$ in (\ref{eq:amplitude1}) on gets \footnote{This yields the exact result as if computed directly from the diagram of Fig.~1.} \begin{equation} _{out}\langle e^+ \nu_e \vert \pi^+\rangle _{in} =\frac{G_F}{\sqrt{2}}\ \langle e^+ \nu_e \vert L_\mu\vert 0\rangle \langle 0 \vert v\ \partial^\mu G^- \vert \pi^+ \rangle. \label{eq:amplitude2} \end{equation} In analogy with (\ref{eq:projection}), one has (still in the case of two generations) \begin{equation} G^+ = {\textswab b}\ (G^+_{ud}\ \pi^+ + G^+_{us}\ K^+ + G^+_{cd}\ D^+ + G^+_{cs}\ D_s^+), \label{eq:goldstone} \end{equation} and, from the relation \begin{equation} \pi^+ = \frac{1}{2{\textswab b}}\left( c_\theta(P^+({\mathbb D}_1) + P^+({\mathbb D}_2)) - s_\theta(P^+({\mathbb D}_3) + P^+({\mathbb D}_4)) \right) \label{eq:pi+1} \end{equation} the inversion of the rotation (\ref{eq:rotation}) defining the $(H,\vec G)$ multiplet (\ref{eq:defHG}) yields \begin{equation} \pi^+ = \frac{1}{2{\textswab b}}\left( (c_\theta (c_1 - s_1c_3) + s_\theta s_1s_3) G^+ + \ldots \right), \label{eq:pi+2} \end{equation} and thus \begin{equation} _{out}\langle e^+ \nu_e \vert \pi^+\rangle _{in} = \frac{1}{2{\textswab b}} \left(c_\theta(c_1 - s_1c_3) + s_\theta s_1s_3)\right) \frac{G_F}{\sqrt{2}}\ \langle e^+ \nu_e \vert L_\mu\vert 0\rangle \langle 0 \vert v\ \partial^\mu G^- \vert G^+\rangle; \label{eq:amplitude4} \end{equation} as the goldstones have been normalized to ``$1$'' \begin{equation} \langle 0 \vert G^- \vert G^+\rangle =1 \end{equation} one obtains \begin{equation} _{out}\langle e^+ \nu_e \vert \pi^+\rangle _{in} = i\, v\, k^\mu\;\frac{1}{2{\textswab b}} \left(c_\theta (c_1 - s_1c_3) + s_\theta s_1s_3\right) \frac{G_F}{\sqrt{2}}\ \langle e^+ \nu_e \vert L_\mu\vert 0\rangle, \label{eq:amplitude5} \end{equation} where $k^\mu$ is the momentum of the incoming pion. When $\textswab b$ is given by (\ref{eq:b}), one recovers the standard PCAC result for \begin{equation} f_\pi = \left((c_1 - s_1c_3) + \frac{s_\theta}{c_\theta}s_1s_3\right) f_0 = \frac{G^+_{ud}}{c_\theta} f_0 . \label{eq:fpi} \end{equation} It is easy to find along the same way \begin{eqnarray} f_K &=& \left((c_1 - s_1c_3) - \frac{c_\theta}{s_\theta}s_1s_3\right) f_0 = \frac{G^+_{us}}{s_\theta} f_0,\cr f_D &=& \left((c_1 + s_1c_3) + \frac{c_\theta}{s_\theta}s_1s_3\right) f_0 = \frac{G^+_{cd}}{(-s_\theta)} f_0,\cr f_{D_s} &=& \left((c_1 + s_1c_3) - \frac{s_\theta}{c_\theta}s_1s_3\right) f_0 = \frac{G^+_{cs}}{c_\theta} f_0. \label{eq:f's} \end{eqnarray} In the limit $\theta_1 = \theta_3=0$ all $f$'s become identical. Note the computation of the decay rate does not introduce extra $\textswab b$ factors in the phase space integral because the outgoing states are leptons, which are normalized to ``$1$''. \subsection{The hierarchy of quark condensates} \label{subsec:condensates} I show now in a precise example how the orientation of the Higgs in flavour space also determines the hierarchy of quark condensates. By our choice of a unique Higgs boson identified with the scalar entry of $\tilde \varphi_1$, we have imposed that it is the only scalar with a non-vanishing vacuum expectation value (VEV). This means a departure from the symmetric case where all diagonal quark condensates have the same VEV and where all non diagonal condensates vanish. The system of equations to be satisfied by the different VEV's is now:\l - for the scalar singlets of the $({\mathbb S}, \vec{\mathbb P})$ multiplets: \begin{equation} \langle \tilde{\mathbb S}_2\rangle = \langle \tilde{\mathbb S}_3\rangle = \langle{\mathbb S}^0({\mathbb D}_4)\rangle = 0; \end{equation} - for the neutral scalars of the $({\mathbb P}^0, \vec{\mathbb S})$ multiplets: \begin{equation} \langle {\mathbb S}^3({\mathbb D}_1)\rangle = \langle {\mathbb S}^3({\mathbb D}_2)\rangle = \langle {\mathbb S}^3({\mathbb D}_3)\rangle = \langle {\mathbb S}^3({\mathbb D}_4)\rangle =0, \end{equation} where the notation $\langle {\mathbb S} \rangle$ used above is a shortcut for $\langle \overline\Psi {\mathbb S} \Psi\rangle$. One case for which these equations can be solved approximately is for example $s_3 \approx 0, c_3 \approx 1$; one finds \vbox{ \begin{eqnarray} & & \langle \bar c c \rangle \approx \frac{c_1 + s_1}{c_1 - s_1}\langle \bar u u \rangle, \cr & & \langle \bar d d \rangle \approx \frac{c_1 -s_1(c_\theta^2 - s_\theta^2)}{c_1-s_1} \langle \bar u u \rangle, \cr & & \langle \bar s s \rangle \approx \frac{c_1 +s_1(c_\theta^2 - s_\theta^2)}{c_1-s_1} \langle \bar u u \rangle, \cr & & \langle \bar d s \rangle = \langle \bar s d \rangle \approx -2 \frac{s_1 s_\theta c_\theta}{c_1 - s_1} \langle \bar u u\rangle, \cr & & \langle \bar u c \rangle = \langle \bar c u \rangle = 0. \label{eq:consensates} \end{eqnarray} } For $c_1>0\;,\;s_1<0$, one gets accordingly the hierarchy \begin{equation} \langle \bar u u \rangle > \langle \bar d d \rangle > \langle \bar s s \rangle >\langle \bar c c \rangle >\langle \bar d s \rangle =\langle \bar s d \rangle > \langle \bar u c \rangle = \langle \bar c u \rangle = 0; \end{equation} it agrees with the one generally admitted from the Gell-Mann-Oakes-Renner relation; the non-vanishing of non-diagonal quark condensates, which appears in a natural manner here, has been debated in the past when dealing with kaon decays \cite{tadpole}. \subsection{The coupling of the Higgs boson to pseudoscalar mesons} \label{subsec:higgs-mesons} I determine here the coupling of the Higgs boson to pseudoscalar mesons in terms of leptonic decay constants and CKM mixing angles. The ``mexican hat'' potential introduced for $\tilde\varphi_1 = (H,\vec G)$ \begin{equation} V(H,\vec G)= -\frac{\sigma^2}{2} \tilde\varphi_1\otimes\tilde\varphi_1 + \frac{\lambda}{4} (\tilde\varphi_1\otimes\tilde\varphi_1)^{\otimes 2} \label{eq:potential} \end{equation} which triggers the breaking of the electroweak symmetry yields in particular a coupling between the Higgs and the goldstones \begin{equation} {\cal L} \ni -\frac{\lambda}{\sqrt{2}}v\; h\; (2G^+G^- + G^3 G^3). \label{eq:phi4} \end{equation} Let $a,b$ the flavour indices of a pseudoscalar meson ${\cal P}_{ab}$, with mass $M_{ab}$ transforming like the diquark operator $\bar q_a \gamma_5 q_b$ (for example ${\cal P}^+_{ud} = \pi^+$); the goldstone $G^+$ writes, according to (\ref{eq:phinorm},\ref{eq:flavoureigenstates},\ref{eq:bdef}) and in analogy with (\ref{eq:goldstone}) \begin{equation} G^+ = \sum_{a,b} G^+_{ab} \Pi^+_{ab} = {\textswab b} \sum_{a,b} G^+_{ab} {\cal P}^+_{ab}\;. \end{equation} Making use of the equivalent of relations (\ref{eq:fpi},\ref{eq:f's}) in the case of three generations, one gets in particular for two charged outgoing flavour eigenstates ${\Pi}^+_{ab}$ and ${\Pi}^-_{cd}$ the coupling \footnote{The Lagrangian is always written with fields which are normalized to ``$1$'', and the corresponding couplings are used to compute $S$-matrix elements between states which are also normalized to ``$1$''. The physical $S$-matrix elements involving asymptotic mesons that we need to compute are then deduced by introducing the appropriate $\textswab b$ factors.} \begin{equation} -i\, \sqrt{2}\lambda\,v\;\sum_{a,b}\sum_{c,d} \frac{f_{ab}f_{cd}}{f_0^2}V_{ab}V^\dagger_{cd} \;h\,{\Pi}^+_{ab}\,{\Pi}^-_{cd} \label{eq:h2mesons} \end{equation} where $f_{ab}$ is the leptonic decay constant of ${\cal P}_{ab}$ and $V_{ab}$ the corresponding entry of the CKM mixing matrix. A dominant feature in (\ref{eq:h2mesons}) is the presence of the CKM mixing angles $V_{ab}$, which, unlike what happens at the quark level for Yukawa couplings, can strongly damp the corresponding coupling independently of the mass of the outgoing particles. This is specially the case for $B^\pm$ mesons, since the corresponding $V_{ub}$ lies far from the diagonal in the CKM mixing matrix. I now use the above results to investigate decays of the $Z$ boson which are mediated by the Higgs. \section{Some decays $\boldsymbol{ Z \rightarrow e^+e^-{\cal P}_i^+{\cal P}_j^-}$} \label{section:decays} I shall always deal with the cases when the outgoing leptons are electrons; when they are muons, the results are very much alike, because the relative difference in the available phase space is very small. \subsection{The decay $\boldsymbol{Z \rightarrow e^+e^-\,B^+B^-}$} \label{subsec:BB} There are two types of contributions shown in Figs.~2a,2b. \vbox{ \begin{center} \epsfig{file=Fig2a.ps,height=6truecm,width=8truecm} \vskip .5cm plus 3mm minus 2mm {\em Fig.~2a: The decay $Z \rightarrow e^+e^- B^+B^-$: direct coupling.} \end{center} } \vbox{ \begin{center} \epsfig{file=Fig2b.ps,height=6truecm,width=8truecm} \vskip .5cm plus 3mm minus 2mm {\em Fig.~2b: The decay $Z \rightarrow e^+e^- B^+B^-$: the contribution of the Higgs boson.} \end{center} } That there exists a direct coupling of two gauge bosons to two mesons differs from the standard model with quarks, where the mesons can only originate from the ``hadronization'' of the two quarks coupled to the Higgs boson trough a Yukawa coupling. We need to compute the amplitude ${\cal A}_{Z \rightarrow B^+B^-e^+e^-}$ \begin{equation} {\cal A}_{Z \rightarrow B^+B^-e^+e^-} = {_{out}\langle e^+e^-\,B^+B^-\vert Z_\mu\rangle _{in}} \label{eqq:amp1} \end{equation} where the asymptotic fields $Z_\mu, e^\pm$ are normalized to ``$1$'' while the $B$ mesons are not (see subsection \ref{subsubsec:flavour}). This introduces $\textswab b$ factors according to \begin{equation} {\cal A}_{Z \rightarrow B^+B^-e^+e^-} = \frac{1}{{\textswab b}^2}\; {_{out}\langle e^+e^-\,\Pi^+_{ub}\Pi^-_{bu}\vert Z_\mu\rangle _{in}} \label{eq:amp2} \end{equation} where we have introduced the flavour eigenstates $\Pi_{ab}$ normalized to ``$1$'' which are also used to express the Lagrangian. Eq.~(\ref{eq:amp2}) can be calculated with standard rules. The total coupling of two $Z$ bosons to $\Pi^+_{ub}\Pi^-_{bu}$ reads (including the ``$i$'' coming from $\exp(iS)$) \begin{equation} i\,\frac{g^2}{2c_W^2}\left( 1 + \frac{1}{2}(\frac{f_B}{f_0})^2 V_{ub}V^\dagger_{bu}\frac{M_H^2}{q^2 - M_H^2} \right) \label{eq:ZPP} \end{equation} where the first contribution comes from the direct coupling and the second from the one involving the Higgs boson; $c_W$ is the cosine of the Weinberg angle; $q$ is the Higgs momentum, the mass of which is (see (\ref{eq:potential},\ref{eq:Hvev})) \begin{equation} M_H^2 = \lambda v^2. \label{eq:MH} \end{equation} Note that the mixing angles $V_{ub}V^\dagger_{bu}$ only appear in the Higgs contribution. A second potential source of damping appears in (\ref{eq:ZPP}) since there can be a destructive interference between the direct contribution and the one with the Higgs boson. The decay rate is expressed as a double integral over the square of the Higgs momentum $s = q^2$ and that of the virtual $Z$ momentum $t = (p-q)^2$ : \vbox{ \begin{eqnarray} \Gamma_{Z\rightarrow e^+e^-B^+B^-} &=& \frac{1}{512\sqrt{2}\pi^5}(1-2s_W^2)M_Z^3 G_F^3\cr & &\int_{4M_B^2}^{(M_Z - 2m_e)^2}ds\ \frac{\lambda^{1/2}(s,M_B^2,M_B^2)}{s} \left( 1 + \frac{1}{2}(\frac{f_B}{f_0})^2 V_{ub}V^\dagger_{bu} \frac{M_H^2}{s - M_H^2} \right)^2\cr & &\int_{4m_e^2}^{(M_Z - \sqrt{s})^2}dt\ \frac{\lambda^{1/2}(M_Z^2,s,t)} {(t-M_Z^2)^2} \left( t +\frac{\lambda(M_Z^2,s,t)}{12\,M_Z^2}\right), \label{eq:gammaBB} \end{eqnarray} } where $M_Z$ is the mass of the $Z$ gauge boson, \footnote{With the normalizations chosen here one has $M_W^2 = g^2v^2/8$ and $G_F/\sqrt{2} = g^2/8M_W^2 = 1/v^2$, where $g$ is the $SU(2)_L$ coupling constant.} $M_B$ the mass of the $B^\pm$ pseudoscalar mesons, $m_e$ the mass of the electron \footnote{The mass of the electron has been neglected in the computation of the square of the amplitude.} and $s_W$ the sine of the Weinberg angle; $\lambda(u,v,w)$ is the fully symmetric function \begin{equation} \lambda(u,v,w) = u^2 + v^2 + w^2 -2\,uv -2\,uw -2\,vw; \label{eq:lambda} \end{equation} All $\textswab b$ factors cancel in the decay rate: the $1/{\textswab b}^4$ coming from the amplitude (\ref{eq:amp2}) squared exactly matches the two factors ${\textswab b}^2$ coming from the phase-space measures (\ref{eq:measure2}) for the two outgoing mesons. The kinematical intervals being respectively $s \in [4M_B^2, (M_Z - 2m_e)^2]$ and $t \in [4m_e^2, (M_Z - \sqrt{s})^2]$, for $M_H^2 < (M_Z - 2m_e)^2$, the potential divergence of the decay rate due to the (squared) propagator of the Higgs has to be smoothed out by introducing a width for the latter; we restrict ourselves to the two-mesons channels, as depicted in Fig.~3. \vbox{ \vskip .5cm plus 3mm minus 2mm \begin{center} \epsfig{file=Fig3.ps,height=4truecm,width=8truecm} \vskip .5cm plus 3mm minus 2mm {\em Fig.~3: Introducing a width for the Higgs boson.} \end{center} } The couplings of the Higgs to two charged mesons are the same as the ones used above and studied in subsection \ref{subsec:higgs-mesons}; we shall however neglect in the computation of its width the misalignment of the Higgs with respect to ${\mathbb S}^0({\mathbb D}_1)$, consider all ratios $f_{ab}/f_0 \approx 1$ and only take into account the channels which are not damped by small mixing angles (combinations of $\pi^\pm$ and $D_s^\pm$ in the charged sector). The neutral mesons ($\pi^0, \eta, \eta_c, \eta_b$) are incorporated along the same lines. We thus replace in (\ref{eq:gammaBB}) \vbox{ \begin{eqnarray} & &\left( 1 + \frac{1}{2}(\frac{f_B}{f_0})^2 V_{ub}V^\dagger_{bu}\frac{M_H^2}{(s-M_H^2)} \right)^2 \longrightarrow \cr &=& 1 + \frac {(\frac{f_B}{f_0})^2 V_{ub}V^\dagger_{bu}M_H^2(s-M_H^2) + \left( \frac{1}{2}(\frac{f_B}{f_0})^2 V_{ub}V^\dagger_{bu}\right)^2 M_H^4} {(s-M_H^2)^2 + \left(\frac{G_FM_H^4}{32\pi\sqrt{2}}\right)^2 \left(\sum_{ab}\sum_{cd} V_{ab}V^\dagger_{ba} V_{cd}V^\dagger_{dc} \frac{\lambda^{1/2}(s,M_{ab}^2,M_{cd}^2)}{s}\right)^2}\cr & & \label{eq:largeur} \end{eqnarray} } where the $\sum_{ab} \sum_{cd}$ in (\ref{eq:largeur}) are performed over all above mentioned couples of $J=0$ pseudoscalar mesons with flavour indices $(ab)$ and $(cd)$ and masses $M_{ab}$ and $M_{cd}$; in the charged sector, the $V_{ab}$'s are restricted to the diagonal entries of the CKM mixing matrix, and stand for $1$ in the neutral sector (see (\ref{eq:reps}) with ${\mathbb D} = 1$). To maximize the decay rate, we take $f_B/f_0 \approx 10$ and $V_{ub}$ is chosen at the upper limit of the experimental bounds \cite{TableParticleProperties} \begin{equation} V_{ub} \approx 4\,10^{-3}. \label{eq:Vub} \end{equation} \vbox{ \vskip .5cm plus 3mm minus 2mm \begin{center} \epsfig{file=Fig4.ps,height=14truecm,width=10truecm,angle=-90} \vskip .5cm plus 3mm minus 2mm {\em Fig.~4: $10^4 \times$ the Higgs contribution to the decay rate $Z \rightarrow e^+e^- B^+B^-$.} \vskip .5cm plus 3mm minus 2mm \end{center} } The Higgs contribution, shown in Fig.~4, exhibits a resonance-like behaviour, with a maximum of $\approx 6.6\,10^{-12}\,GeV$ for $M_H \approx 13\;GeV$, which is extremely small \footnote{The numerical integrations have been performed by the method of Newton and dividing, for each point of the graphs, the two dimensional domain of integration into $5\,10^3 \times 10^3$ cells, ensuring a perfect stability.} . The ``background'' coming from the contribution where no Higgs is involved is nearly $1000$ times larger than the Higgs contribution at its maximum and is itself outside the reach of present observations, with a rate \footnote{The destructive interference effects between the two contributions is of course negligeable.} \begin{equation} \Gamma^{no\ Higgs}_{Z\rightarrow e^+e^- B^+B^-} = 5\,10^{-9}\;GeV, \label{eq:BBnoH} \end{equation} to be compared with the total width of the $Z$ boson $\Gamma_Z \approx 2.4\;GeV$; suppose that one can analyze $20\, 10^6$ $Z$ decays \cite{EWmeasures}; ten identified decays into two leptons and two pseudoscalar mesons would correspond to a fraction $5\,10^{-7}$ of all and to a partial width $\Gamma_{part} = 1.2\,10^{-6}\, GeV$; it seems consequently reasonable to set an (optimistic) threshold of observability above a partial width \begin{equation} \Gamma_{obs} \geq 10^{-6}\,GeV. \label{eq:thresh} \end{equation} Accordingly, it appears useless to look for a Higgs boson like the one described above in the decay $Z \rightarrow e^+e^- B^+B^-$. \subsection{The decay $\boldsymbol{Z \rightarrow e^+e^-\,D_s^+D_s^-}$} \label{subsec:FF} I then analyze the decay $Z \rightarrow e^+e^- D_s^+D_s^-$; the corresponding rate is computed from formulae similar to (\ref{eq:gammaBB},\ref{eq:largeur}), with the appropriate substitutions concerning the masses, leptonic decay constants and mixing angles. The damping due to CKM mixing angles is now negligeable, which makes the Higgs contribution dominate over the direct coupling of the $Z$ to two mesons. It appears consequently as a better reaction to look for the Higgs boson. The background coming from from the direct coupling of the $Z$ to two mesons is negligeable: \begin{equation} \Gamma^{no\ Higgs}_{Z\rightarrow e^+e^- D_s^+D_s^-} = 7.6\,10^{-9}\;GeV. \label{eq:FFnoH} \end{equation} The results are shown in Figs.~5. \vbox{ \vskip .5cm plus 3mm minus 2mm \begin{center} \epsfig{file=Fig5a.ps,height=14truecm,width=10truecm,angle=-90} \vskip .5cm plus 3mm minus 2mm {\em Fig.~5a: The decay rate $Z \rightarrow e^+e^- D_s^+D_s^-$.} \end{center} } \vbox{ \vskip .5cm plus 3mm minus 2mm \begin{center} \epsfig{file=Fig5b.ps,height=14truecm,width=10truecm,angle=-90} \vskip .5cm plus 3mm minus 2mm {\em Fig.~5b: The decay rate $Z \rightarrow e^+e^- D_s^+D_s^-$.} \end{center} } The threshold of observability (\ref{eq:thresh}) is satisfied for $M_H \leq 65\,GeV$. This is to be compared with the present lower bound of the four LEP experiments for the Higgs boson of the standard model $M_{H\,standard}^{LEP} \geq 90\,GeV$ at $95\%$ confidence level \cite{HiggsLEP}. In this mass range, there is no hope to detect the Higgs that is proposed here, but it could have been missed at lower masses (but high enough for the missing energy channel to be undetectable \cite{Higgsexp}) because of a too low statistics. \subsection{The decay $\boldsymbol{Z \rightarrow e^+e^-\,K^\pm\pi^\mp}$} \label{subsec:KPI} There also exist in this approach decays which do not occur in the Glashow-Salam-Weinberg model for quarks and characterize a composite Higgs like has been introduced above; they are the ones corresponding to ``flavour changing neutral currents'' in the scalar sector. They could furthermore be easily identified experimentally if produced with enough statistics. {From} (\ref{eq:phi4}) and (\ref{eq:goldstone}) or its generalization to three generations, it appears that the Higgs can decay into final states like $K^\pm \pi^\mp, D^\pm K^\mp \ldots$, and that, due to the choice of the kinetic terms as stated in subsections \ref{subsubsec:lagrangian} and \ref{subsubsec:invariants} and to the property of diagonalization of the corresponding quadratic invariant (\ref{eq:diaginvar}), the background described in the two previous decays is now absent. I study here the $e^+ e^- K^\pm \pi^\mp$ final state, the amplitude of which benefits from a moderate damping by the mixing angles, only $s_\theta c_\theta$. The decay rate is computed form (\ref{eq:gammaBB},\ref{eq:largeur}) where the factor ``$1$'' in each of them, corresponding to the direct coupling, is dropped since the latter does not exist any more. The results for the decay rate are plotted in Fig.~6. \vbox{ \vskip .5cm plus 3mm minus 2mm \begin{center} \epsfig{file=Fig6.ps,height=14truecm,width=10truecm,angle=-90} \vskip .5cm plus 3mm minus 2mm {\em Fig.~6: The decay rate $Z \rightarrow e^+e^- K^\pm \pi^\mp$.} \end{center} } It exhibits the same resonance-like behaviour as the previous decays, but has its maximum $\Gamma_{max} \approx 4.7\,10^{-6}\,GeV$ at a low Higgs mass $M_H \approx 1\,GeV$. The cusps that can be seen on the curve have been checked to correspond to the opening of the different two-mesons channels in the propagator of the Higgs boson. The resonance is rather sharp, and it appears that for $M_H > 7\,GeV$ the width of the process is lower than the threshold (\ref{eq:thresh}) above which there is no hope that it be observed. As a Higgs mass $M_H < 7\,GeV$ would have been detected by the corresponding missing energy \cite{Higgsexp} (with the possible {\it caveat} of footnote \ref{foot:asympt}), there seems unfortunately to be no hope to observe the characteristic decays mentioned above in a foreseeable future. \section{Conclusion} \label{section:conclusion} Answering a demand that alternatives to the strictly standard Higgs boson or to its supersymmetric extensions be searched for. I have proposed one, which, though it shares many similarities with the standard model, exhibits a Higgs boson that is still more elusive; it may not however be beyond experimental reach. It has been incorporated in a framework where the fields in the Lagrangian are tightly related with the $J=0$ mesonic states observed asymptotically; I have related the orientation of the Higgs in flavour space to leptonic decay constants of pseudoscalar mesons and to the CKM mixing angles, such that the quartic potential which triggers the breaking of the electroweak symmetry has been expressed in terms of these parameters and of the mass of the Higgs itself. It includes in particular the coupling of the Higgs to two pseudoscalar mesons, which is no longer triggered by Yukawa couplings to quarks followed by an hadronization process as it used to be in the standard model. New links have thus been provided, which enabled the study of the disintegration of the $Z$ gauge boson into two leptons and two pseudoscalar mesons. I showed that this Higgs boson has interesting and specific properties, in particular that it can trigger flavour changing neutral currents, by decaying into final states of the type $K^\pm \pi^\mp$; unfortunately, detecting those decays would require a tremendous increase of the available number of $Z$ bosons. The decay $Z \rightarrow e^+e^- D_s^+ D_s^-$ is the best candidate because it is not damped by small mixing angles and the background due to the direct coupling of the $Z$ to two mesons is negligeable. A Higgs with mass lower that $65\,GeV$ could have been missed. Unlike in the standard model for quarks, the final state $e^+e^- B^+B^-$ appears to be undetectable because of the presence of the CKM mixing angles in the Higgs coupling to two pseudoscalar mesons; the Higgs contribution is furthermore screened by a background which, though much larger, is itself undetectable. I did not explicitly present here the results for another channel which is not suppressed by small mixing angles, $Z \rightarrow e^+ e^- \pi^+ \pi^-$: the reason is that the corresponding decay rate is then peaked at very low Higgs masses $M_H < 1\,GeV$, for which such a particle could not have escaped detection in the missing energy channel; for higher masses it becomes again absolutely undetectable. For large Higgs masses ($M_H > M_W$), the type of decays studied here is undetectable, and more standard considerations have to be pursued. \vskip 1cm \begin{em} \underline {Acknowledgments}: it is a pleasure to thank F. Boudjema for critics and suggestions. \end{em} \newpage\null \listoffigures \bigskip \begin{em} Fig.~1: Leptonic decay of a pseudoscalar meson;\l Figs.~2: The two contributions to the decay $Z \rightarrow e^+e^-\; B^+B^-$;\l Fig.~3: Introducing a width for the Higgs boson;\l Fig.~4: $10^4 \times$ the Higgs contribution to the rate of the decay $\Gamma_{Z \rightarrow e^+e^-\, B^+B^-}$ as a function of the Higgs mass;\l Figs.~5: The decay rate $\Gamma_{Z \rightarrow e^+e^-\, D_s^+D_s^-}$ as a function of the Higgs mass;\l Fig.~6: The decay rate $\Gamma_{Z \rightarrow e^+e^-\, \pi^+K^-}$ as a function of the Higgs mass. \end{em} \newpage\null {\Large\bf Appendix}
\section{Introduction} Structure forms in hierarchically clustering universes as primordial dark matter density fluctuations are amplified by gravity and collapse in a constantly evolving population of virialized dark matter halos. In this scenario galaxies are envisioned to form as baryons follow the dark matter collapse, dissipate their kinetic energy through shocks, and radiate it away as they settle (and form stars) in centrifugally supported structures at the center of dark halos. Galaxies evolve afterwards as a result of mergers between protogalaxies and of further accretion of intergalactic gas (White \& Rees 1978, Navarro \& White 1993, Cen \& Ostriker 1993, Navarro, Frenk \& White 1994, Evrard, Summers \& Davis 1994, Katz, Weinberg \& Hernquist 1996, Bryan et al 1998, Couchman, Thomas, \& Pearce 1995, Yepes et al 1997, Navarro \& Steinmetz 1997, Tissera, Lambas \& Abadi 1997, Steinmetz \& Navarro 1998). Gravity, pressure gradients, hydrodynamical shocks, and the ability of gas to radiate are therefore physical processes of crucial importance during the formation of galaxies in a cosmological context. Numerical experiments intended to simulate galaxy formation must therefore capture accurately these essential ingredients on the many different levels of the hierarchy that coexist at a given time. Unfortunately, detailed analytic solutions are not known for relevant analogues of the complex galaxy formation process and it has been difficult to assess properly the accuracy and reliability of these codes (Frenk et al 1999). Previous studies have therefore focussed on the sensitivity and convergence of the results regarding numerical parameters such as the size of the grid used in Eulerian hydrodynamical methods (Cen 1992, Bryan et al 1998) or the number of particles used in particle-based methods such as the Smooth Particle Hydrodynamics (SPH, see Gingold \& Monaghan 1977, Benz 1990, Hernquist \& Katz 1989, Navarro \& White 1993, Summers 1993 for general introductions to SPH). Convergence as resolution improves is a necessary condition for simulations to be reliable, but is often not sufficient to ensure that the results are accurate and robust. There is clearly a need for analytic solutions that describe physical situations similar to the galaxy formation scenario envisioned in cosmological models and that can be used to gauge the performance of cosmological hydrodynamical codes. Spherical infall is one relevant situation for which a detailed solution is known. Bertschinger (1985) first computed the detailed behaviour of collisional gas being accreted onto a point mass perturber in an Einstein-de Sitter universe. Assuming that only gravity, pressure gradients, and hydrodynamical shocks control the gas behaviour, Bertschinger exploited the scalefree nature of all these processes to derive similarity solutions that offer a useful testbed for hydrodynamical codes (Navarro \& White 1993, Summers 1993). One crucial ingredient of the galaxy formation process is, however, missing from these tests: radiative cooling. This is because the cooling function is the result of atomic processes that are not independent of scale and therefore similarity solutions of the spherical infall problem are not generally available when radiative energy losses are included. This is true even in the very simplified case when radiation transfer, heat conduction, and magnetic effects are neglected. Self-similarity may still be recovered in situations that include radiative cooling at the expense of placing restrictions on the temperature and density dependence of the cooling function. For example, Bertschinger (1989) computed the detailed self-similar evolution of cooling flows in isothermal potentials under the assumption that the gas cooling function is a power law of density and temperature, $\Lambda(\rho,T) \propto \rho^2 T^{\lambda}$. For $\lambda=1/2$, this power law resembles the contribution from thermal bremsstrahlung to the overall cooling function, so these results can be usefully applied to the hot, diffuse X-ray emitting gas that fills the intracluster medium of rich galaxy clusters. Unfortunately, these solutions are only valid for isolated systems originally in hydrostatic equilibrium and therefore their applicability to problems where continuous mass accretion play a significant role is limited. A related approach has been recently described by Owen, Weinberg \& Villumsen (1988), who derive a family of cooling functions that ensure self-similar evolution in Einstein-de Sitter universes with power-law initial density fluctuations. In this case, similarity is preserved by ensuring that the cooling timescale of an object with characteristic clustering mass ($M_{\star}$) is a fixed fraction of the Hubble time. In this paper we follow a similar approach and derive similarity solutions for the spherical infall problem that include energy losses due to radiative cooling. Similarity is preserved by choosing a convenient power-law form of the cooling function, and its solutions are compared with the results of direct numerical simulations using a hydrodynamical SPH code. Because of the restrictions placed on the cooling function, the applicability of the results to realistic models of galaxy formation is not straightforward, but the solutions are very useful as tests of hydrodynamical codes under physical conditions that combine the major ingredients of galaxy formation models: gravitational collapse, pressure gradients, energy dissipation through shocks, and radiative energy losses, {\it in a proper cosmological context.} We derive the similarity solutions in \S2 and compare them with numerical simulations in \S3. Section 4 discusses the results and \S5 summarizes our main conclusions. \begin{figure*} \centerline{\psfig{file=fig1.ps,width=14cm}} \caption{ Similarity solutions for different choices of the shape of the initial density perturbation. The three curves show the solutions for $\epsilon=1$ (solid lines), $2/3$ (dotted lines), and $1/3$ (dashed lines), respectively. Symbols are as defined in eqs.(3). These analytic solutions neglect radiative energy losses.} \label{fig:nocool} \end{figure*} \begin{figure*} \centerline{\psfig{file=fig2.ps,width=14.0cm}} \caption{Examples of the three types of similarity solutions found when radiative cooling is included, computed for the case $\epsilon=1$ and $K_0=0.1$. The stagnation, adiabatic, and eigensolution correspond to the dotted ($\lambda_s=0.23$), dashed ($\lambda_s=0.15$), and solid line ($\lambda_s=0.1869$), respectively. The eigensolution represents a limiting case of the two other kinds, when the stagnation point approaches the center and the flow extends all the way to $\lambda=0$. } \label{fig:lambdas} \end{figure*} \begin{figure*} \centerline{\psfig{file=fig3.ps,width=14cm}} \caption{The density, velocity, temperature, and enclosed mass profiles showing the effect of increasing the relative importance of cooling. All curves correspond to the ``eigensolution'' for $\epsilon=1$. Shock radii are given in Table 1. The solid, dotted, and dashed curves correspond to $K_0 = 0.0$, $0.1$ and $0.3$, respectively.} \label{fig:k0plot} \end{figure*} \begin{figure*} \centerline{\psfig{file=nocool.ps,width=14cm}} \caption{ Comparison between similarity solutions and the results of SPH simulations. The solution without cooling is represented by the solid line. We also show, for comparison, the solution including radiative energy losses ($K_0=0.1$). Different symbols correspond to the SPH simulation at different times. Open triangles, squares and filled circles correspond to times when $\sim 7$, $15$, and $20\%$ of the initial mass lies inside the shock radius. The results of the simulations are seen to converge to the analytic solution at later times, as more particles pass through the shock and the effects of numerical resolution become less important.} \label{fig:nocool} \end{figure*} \begin{figure*} \centerline{\psfig{file=cool.ps,width=14cm}} \caption{ As in Figure 4. The SPH simulation results shown now include cooling, $K_0=0.1$. The times shown are analogous to those chosen in Figure 4. The SPH simulation reproduces the analytical results very well inside the shock radius. } \label{fig:cool} \end{figure*} \section{Accretion of Collisional Gas onto Scale Free Perturbations} \subsection{Similarity solutions neglecting radiative cooling} The evolution of a density perturbation in an Einstein-de Sitter universe is expected to be self-similar if the initial perturbation contains no physical scales. Following Fillmore \& Goldreich (1984) we characterize the initial perturbation at some initial time $t_i$ by the excess mass within a radial shell of (initial) radius $r_i$, $$ {\delta M_i \over M_i} = \left({M_i \over M_0}\right)^{-\epsilon}, \eqno(1) $$ where $M_i=(4/3)\pi \rho_{H} r_i^3$ is the unperturbed mass within $r_i$, $\rho_{H}=3H_i^2/8\pi G$ is the critical density for closure, $H_i$ is Hubble's constant at $t=t_i$, $M_0$ is some reference mass, and $\epsilon>0$. Because the mass excess is positive each radial shell is bound to the center and collapses after reaching a (maximum) turnaround radius, $r_{ta}$. The turnaround radius increases with time as $$ r_{ta} \propto t^\xi, \eqno(2) $$ where $\xi=(2/3)(1+1/3\epsilon)$. The mass inside the turnaround radius then grows as $M_{ta}=M(r<r_{ta}) \propto t^{2/3\epsilon} \propto (1+z)^{-1/\epsilon}$, where $z$ is the usual definition of redshift. This scaling can be compared with that of the characteristic clustering mass in a scalefree hierarchical clustering universe where the power spectrum of initial density fluctuations is $P(k) \propto k^{n}$: $M_{\star}(z) \propto (1+z)^{-6/(n+3)}$. Perturbations characterized by a given value of $\epsilon$ therefore accrete mass at the same rate as a ``typical'' mass concentration in a scalefree universe with $n=3(2\epsilon-1)$. A central point mass perturbation corresponds to $\epsilon=1$, with $r_{ta}\propto t^{8/9}$ and $M_{ta}\propto t^{2/3} \propto (1+z)^{-1}$. This is the case considered by Bertschinger (1985). Assuming that at $t=t_i$ the initial velocity field is pure unperturbed Hubble flow, $v=H_i r=(2/3)r/t_i$, there are no further scales in the problem once the magnitude of initial density perturbation has been specified, and the time evolution of the system must approach self-similarity after a short initial transient. This implies that a unique solution, expressed in properly scaled variables, describes the structure of system at all times $t \gg t_i$. It is convenient to express this solution in nondimensional form and, following Bertschinger (1985), we define dimensionless radii, velocities, densities, pressures, masses, and temperatures as follows, $$ \lambda(r,t)={r\over r_{ta}} \eqno(3.1) $$ $$ v(r,t)={r_{ta}\over t} V(\lambda) \eqno(3.2) $$ $$ \rho(r,t)=\rho_H(t) D(\lambda) \eqno(3.3) $$ $$ p(r,t)=\rho_H(t) \left({r_{ta} \over t}\right)^2 P(\lambda) \eqno(3.4) $$ $$ m(r,t)={4\pi \over 3} \rho_H r_{ta}^3 M(\lambda) \eqno(3.5) $$ $$ T(\lambda)={P(\lambda) \over (\gamma-1) D(\lambda)}, \eqno(3.6) $$ where $\gamma=5/3$ is the usual ratio of specific heats. The equations describing the motion of a collisional fluid with spherical symmetry can be expressed in terms of these dimensionless variables and are given by (see Bertschinger 1985 for details), $$ (V-\xi\lambda)D' + DV' + {2DV\over\lambda} -2D = 0 \eqno(4.1) $$ $$ (V-\xi\lambda)V'-(1-\xi)V=-{P'\over D} - {2\over9}{M\over\lambda^2} \eqno(4.2) $$ $$ (V-\xi\lambda)\left({P'\over P}-\gamma{D'\over D}\right) = 2(2-\xi) - 2\gamma \eqno(4.3) $$ $$ M' = 3\lambda^2 D\eqno(4.4) $$ Here primes refer to differentiation relative to $\lambda$. Eqs.(4) are, respectively, the continuity, Euler, adiabatic, and mass equations and are valid for a pressurized fluid flow neglecting radiation, heat conduction, and deviations from spherical symmetry. Pressurization of each radial shell of fluid occurs soon after turnaround as the collapsing shell encounters previously collapsed ones. Because of similarity constraints, the radius at which the shock occurs must be a constant fraction of the turnaround radius, $\lambda=\lambda_s$. Outside $\lambda_s$ the evolution of the gas is identical to the turnaround and collapse of a pressureless shell of material. A full solution of eqs.(4) can be found by locating the radius of the shock and applying Hugoniot shock jump conditions to the exterior pressureless infall values. A parametric form of the preshock cold accretion flow is given by, $$ \lambda = \sin^2(\theta/2) \,\left(\theta-\sin\theta\over\pi\right)^{-\xi} \eqno(5.1) $$ $$ V(\lambda) = \lambda {\sin\theta(\theta-\sin\theta)\over(1-\cos\theta)^2} \eqno(5.2) $$ $$ D(\lambda) ={9\over2}{(\theta-\sin\theta)^2\over(1-\cos\theta)^3(1+3\epsilon\chi)} \eqno(5.3) $$ $$ M(\lambda) = \lambda^3{9\over2}{(\theta-\sin\theta)^2\over(1-\cos\theta)^3} \eqno(5.4) $$ with $\chi = 1 - (3/2)(V(\lambda)/\lambda)$. The shock location depends also on the central boundary conditions, which we take to be that the velocity and mass must vanish, i.e. $V=M=0$ at $\lambda=0$. Identifying the values of the variables inside (outside) the shock by the subscript 2 (1), we have, at $\lambda=\lambda_s$, $$ {(V_2-\xi\lambda_s)\over(V_1-\xi\lambda_s)}={\gamma-1 \over \gamma+1} \eqno(6.1) $$ $$ D_2=\left(\gamma-1 \over \gamma+1\right) D_1 \eqno(6.2) $$ $$ P_2={2\over\gamma+1}D_1(V_1-\xi\lambda_s)^2 \eqno(6.3) $$ $$ M_2=M_1. \eqno(6.4) $$ Figure 1 shows the resulting density, velocity, temperature and entropy profiles for various values of the initial perturbation parameter $\epsilon$. As $\epsilon$ decreases from unity (the value corresponding to a point mass perturbation, see solid line) to $1/3$ (dashed line), the shock moves inwards, the inner density profile becomes shallower and the temperature profile becomes approximately isothermal. We shall see next how these results are altered by the inclusion of radiative cooling effects. \subsection{Similarity solutions including radiative energy losses} \subsubsection{The self-similar cooling function} The results discussed in the previous subsection are only applicable in the limiting case when energy losses due to radiative cooling are neglected. As discussed in \S1, the cooling function of a plasma with realistic cosmic abundances has a complex dependence on temperature and imposes dimensional physical scales on the problem that violate the conditions required for the existence of self-similar solutions. Similarity solutions may exist only when the cooling processes introduce no further scales in the problem. This condition can be satisfied by choosing an appropriate cooling function so that the overall cooling efficiency is independent of time. This can be ensured by demanding, for example, that the cooling radius (i.e. the radius at which the cooling time equals the age of the universe) be a fixed fraction of the turnaround radius of the system. Equivalently, one may require that, at some fixed fraction of the turnaround radius, the ratio between the local cooling timescale and the age of the universe be constant and independent of time. The cooling time is given by $$ t_{cool}={u\over du/dt}={u \rho \over \Lambda(\rho,T)} \propto {u \rho_H \over \Lambda(\rho,T)}, \eqno(7) $$ where $u\propto T$ is the specific thermal energy of the gas, and the proportionality in eq.(7) is valid at a fixed value of $\lambda$. The condition $$ t_{cool}/t_H=(6\pi G\rho_H)^{1/2} t_{cool}=K_0^{-1}={\rm constant} \eqno(8) $$ is thus satisfied if $$ \Lambda(\rho,T) \propto \rho^{3/2} u \propto \rho^{3/2} T. \eqno(9) $$ This condition is independent of $\epsilon$ and implies that the solution will be self similar regardless of the time dependence of the turnaround radius. Our similarity solutions thus require a weaker dependence on density and a stronger dependence on temperature than expected from thermal bremsstrahlung emission, $\Lambda_{bremss} \propto \rho^2 T^{1/2}$. We note, however, that eq.(9) is not the only cooling function that would lead to self-similar evolution. In particular, since the characteristic ``virial temperature'' ($T_{vir} \propto GM_{ta}/r_{ta}$) of the system is related to its mean density by the growth rate of the turnaround radius, it is possible to retain the $\rho^2$ dependence characteristic of realistic cooling functions and adjust only the temperature exponent to preserve similarity. The price one pays is that in this case the temperature exponent of the self-similar cooling function depends on $\epsilon$. More explicitly, $\Lambda(\rho,T) \propto \rho^2 T^{\beta}$, with $\beta=1-(9/2)(\epsilon/(3\epsilon-2))$ (Owen et al 1998). In this case, the relative velocities of the cooling radius and the shock radius are equal. We emphasize that our choice of self-similar cooling function (eq.9) is independent of $\epsilon$ and does not rely on tuning the two velocities to agree with each other. \subsubsection{The similarity solutions} Once the appropriate form of the cooling function has been chosen, the behaviour of the gas can be computed using eqs.(4) after modifying the entropy conservation equation (4.3) to allow for energy losses. In dimensionless form, the modified eq.(4.3) now reads $$ (V-\xi\lambda)\left({P'\over P}-\gamma{D'\over D}\right) = 2(2-\xi) - 2\gamma -K_0 D^{1/2}. \eqno(10) $$ At any radius inside the shock, and at all times, the ratio between the local dynamical time ($3\pi/16G\rho)^{1/2}$ and the cooling time equals $\pi \sqrt{18/16} K_0$. Equations 4.1, 4.2, 4.4, and 10 can now be solved to describe the post-shock flow once adequate boundary conditions at the center are imposed. As discussed by Bertschinger (1989), three different kinds of solutions can be identified according to the limiting behaviour of the solution near the center. The first type is a solution where the flow stagnates at some finite radius. Infalling gas settles onto this surface, where the density formally diverges. In order to obey self-similarity the surface must move outwards at the same rate as the turnaround radius. This kind of solution thus requires a piston to move the surface outwards and is therefore of little physical applicability. The second type of inner solution corresponds to a flow that extends all the way to $\lambda=0$ so that the local flow time, $t_{flow}=r/v$, near the center becomes much shorter than the cooling time. These solutions are referred to as ``adiabatic'' solutions, because cooling is unimportant for small $\lambda$. The mass accretion rate near the center approaches a constant and the accretion speed diverges near the center. The third kind of solution, sometimes called the ``eigensolution'', is the limiting adiabatic solution with minimum central mass accretion rate, or, equivalently, the limit of the family of stagnating solutions as the stagnation radius tends to zero. Each of these solutions is characterized by different values of the shock radius, $\lambda_s$. Values of $\lambda_s$ similar to those obtained neglecting cooling correspond to solutions with stagnation points. As the shock radius moves inwards the stagnation point moves closer to the center and the solution transitions through the eigensolution to the adiabatic case. Figure 2 shows examples of these three different solutions for the case $\epsilon=1$, $K_0=0.1$. The solution with $\lambda_s=0.23$ (dotted line) has a stagnation point at $\lambda_0 \sim 0.026$ where the density diverges and the velocity becomes zero at the surface. There is no mass inside this radius, and the surface is pushed out by a piston to preserve similarity. As the shock radius is reduced to $\lambda_s=0.15$ the flow extends all the way to the center and the infall velocity diverges there. The eigensolution corresponds to $\lambda_s \approx 0.1869$. In this case the central velocity remains finite at the origin and the flow extends all the way to the center. In practice, we find this solution numerically by letting the stagnation point, $\lambda_0$, approach zero. \begin{table} \begin{center} \begin{tabular}{l|l|l|l} \hline $\epsilon$ & $\xi$ & $K_0$ & $\lambda_s$\\ \hline 1 & $8/9$ & 0.0 & 0.3389\\ 1 & $8/9$ & 0.1 & 0.1858\\ 1 & $8/9$ & 0.3 & 0.0939\\ \\ 2/3 & 1 & 0.0 & 0.2899\\ 2/3 & 1 & 0.1 & 0.1551\\ 2/3 & 1 & 0.3 & 0.0733\\ \\ 1/3 & $4/3$ & 0.0 & 0.1889\\ 1/3 & $4/3$ & 0.1 & 0.0996\\ 1/3 & $4/3$ & 0.3 & 0.0422\\ \hline \end{tabular} \caption{The dimensionless shock radius of the eigensolution for various choices of the shape parameter of the initial density perturbation, $\epsilon$, and of the dimensionless cooling coefficient, $K_0$. The time exponent of the turnaround radius, $\xi$, is also listed for each case.} \label{tab:solns} \end{center} \end{table} Noting that the solutions have approximately constant infall velocity in the inner regions, it is possible to determine the asymptotic slopes of the density and pressure profiles. As the velocity tends to a constant, $D(\lambda)\rightarrow D_0 \lambda^{-2}$ and $P(\lambda)\rightarrow P_0\lambda^{-3}$, where the normalising constants depend on the cooling parameter $K_0$. The asymptotic velocity is given by $- K_0 D_0^{1/2} / (2\gamma-3)$. Figure 3 shows how the eigensolutions vary as a function of the cooling efficiency parameter $K_0$. As the importance of cooling increases the pressure support inside the shock decreases and the shock radius moves inwards. Perhaps counterintuitively, as cooling becomes more important the temperature inside the shock radius {\it increases} and the density {\it decreases}. This is because low entropy gas is ``lost'' to the central mass and, at fixed radius, low entropy gas is replaced by higher entropy gas that moves in from outside. The ``cooling flow'' thus results in a net {\it increase} in gas entropy at a given radius. Figure 3 illustrates that cooling has a substantial effect on the structure of the system, and suggests that similarity solutions with cooling may provide a stringent test of the capabilities and accuracy of hydrodynamical codes. We pursue this issue next. \section{Comparison with SPH simulations} As discussed in \S1, the solutions derived in the previous section can be fruitfully confronted with the results of cosmological hydrodynamical codes. This comparison is all the more interesting because the test case we discuss in the previous section captures many of the salient features of the galaxy formation process: gravitational collapse, pressurization through shocks, radiative energy losses, cooling flows. Furthermore, because the solutions are self similar in time, a single simulation can be examined at different times and convergence can be directly assessed. This is important because the importance of numerical resolution varies with time within a single simulation. For example, the number of particles within the shock radius increases with time, and the ratio between the smallest resolved radius and the turnaround radius decreases with time. Because the solution is unique, analyzing the deviations between analytic solution and numerical experiment at different times provides invaluable insight into the role of numerical limitations and their consequence on the subsequent evolution of the system. We use the Smooth Particle Hydrodynamics code described by Navarro \& White (1993), where details about the numerical procedure should be consulted. The initial setup is also similar to that described by these authors. We simulate a spherical region of an Einstein-de Sitter universe by laying down $24,257$ particles homogeneously inside a sphere of radius $R$. Each particle is given an initial velocity consistent with unperturbed Hubble flow, and an external potential is added to mimic a point mass perturbation of mass equal to $5\%$ of the total mass of the sphere. The external potential is ``softened'' inside a fixed radius $R_p=0.1R$ in order to prevent divergences, but is fully Keplerian outside $R_p$. This corresponds to the case $\epsilon=1$ in eq.(1). The gravitational softening of each particle is also chosen to be equal to $R_p$. All particles have initially the same temperature, chosen to be much lower than the final virial temperature of the system in order to prevent hydrodynamical effects from becoming important before the gas turns around and passes through the shock. Two different simulations were performed, one with $K_0=0$ and one with $K_0=0.1$. The simulations are evolved until the turnaround radius encompasses half of the total number of particles. At the final time, about $25\%$ of the particles have passed through the accretion shock. Figures 4 and 5 show the dimensionless profiles of density, velocity, and temperature averaged in spherical bins of constant logarithmic width. The dimensionless mass enclosed inside each bin is also shown in the bottom right panel. Each panel shows the result of the simulations at three different times, corresponding to different numbers of particles within the shock radius: $1,838$ (triangles), $3,648$ (squares), and $5,443$ (circles) for the simulation without cooling and $1,655$ (triangles), $3,371$ (squares), and $4,734$ (circles) for the simulation with cooling. The analytic eigensolutions corresponding to $K_0=0$ and $K_0=0.1$ are shown with a solid and dashed line, respectively. The simulation without cooling is in all respects similar to that reported by Navarro \& White (1993), except for the fact that we use about ten times more particles. As shown in Figure 4, the density and velocity profiles agree remarkably well with the analytic solution inside the shock radius $\lambda_s$. As expected, agreement with the analytic solution improves as more and more particles pass through the shock and the post-shock flow becomes better resolved. This is especially noticeable in the mass panel, where it is seen that at early times (triangles) the simulation deviates from the analytic solution but that the system converges to the right solution at later times. Once about $5,000$ particles have gone through the shock the agreement between the solution and the experiment is remarkably good. The convergence towards the similarity solution is also convincingly demonstrated in the temperature panel, where the linear temperature scale accentuates the discrepancies near the center. One important conclusion from this analysis is that although numerical limitations compromise the results of the numerical simulations at early times, these discrepancies have no major effect on the behaviour of the system at late times. The main deviations from the analytic solutions actually happen beyond the nominal shock radius. The shock is smoothed over several resolution lengths, and even at late times only outside approximately $2 \lambda_s$ the preshock flow solution is recovered in the simulations. As discussed in \S2.1 and illustrated directly by the dashed and solid lines in Figures 4 and 5, the similarity solution changes substantially when radiative cooling is included. The shock radius moves inwards, the post-shock temperature increases, the infall velocity is non zero all the way to the center, and the density decreases at all radii because of the accumulation of cold material at the center. Figure 5 shows that the numerical results match very well these predicted changes in the post-shock region. Inside $\lambda_s=0.1872$ the temperature, mass, density, and velocity profiles are almost indistinguishable from the similarity solutions. Remarkably, this is the case even when only $7\%$ ($\approx 1,600$ particles) of the mass of the system has passed through the shock, and the agreement is seen to improve as more and more particles go through the shock. As noted above, the main shortcoming of the simulation regards the width of the shocked region: particles are seen to respond to the shock as far out as $\sim 3$ times the nominal shock radius. It is encouraging, however, to note that the overall trend is correct, and that the volume of the shocked region is substantially smaller than in the case where cooling is neglected. \section{Discussion} One important goal of the comparison between simulations and similarity solutions presented above is to assess the reliability of numerical techniques currently being used to simulate the formation of galaxy-sized structures in the universe. The present study indicates that, to a large degree, SPH methods give results that are consistent with analytic solutions in test cases that involve some of the major ingredients believed to play a significant role during galaxy formation. In particular, our results show that SPH codes can reproduce faithfully the changes in temperature, density, and velocity that are associated with strong ``cooling flows'' onto a central perturbation. The mass inflow rates are also accurately reproduced, and there is no indication that numerical limitations lead to an undue increase or decrement in the amount of mass that cools and flows to the central object. This is important because it has been argued that SPH-like treatments of numerical hydrodynamics may cause an artificial ``overcooling instability'' that exaggerates the importance of radiative cooling losses and leads to the formation of excessively massive gas concentrations at the center of dark halos (Thacker et al 1998). Our study shows that this instability is not present in the test cases we present here. The rate at which gas cools and gets accreted onto a central clump is proportional to the mass accretion rate through the shock radius and is consistent with that expected from the similarity solution. There are, however, important differences between the cases considered in our study and in that of Thacker et al. These authors consider gas cooling from a hot gaseous halo {\it in hydrostatic equilibrium} within a dark halo and report a strong dependence of the cooled mass on the numerical resolution of SPH simulations. They also consider a more realistic cooling function that varies with temperature and density in very different ways than the self-similar cooling function we adopt here. Because of these caveats, it may be premature to argue either for or against their findings on the basis of the simulations presented here. It should be possible, however, to test their arguments using the cooling wave similarity solutions derived by Bertschinger (1989). These are a much closer analogue to the case considered by Thacker et al and direct comparison between numerical experiments and Bertschinger's analytic solutions should provide a definitive assessment regarding the effects of numerical resolution on the cooling and accretion of gas at the center of dark halos. We plan to carry out this comparison in the near future. \section{Summary} We have derived similarity solutions that describe the spherical collapse, shock, and settling of collisional gas from scalefree perturbations in an Einstein-de Sitter universe. Our study extends prior work on the subject by taking into account the full effects of energy loss due to radiative cooling processes, under the simplifying assumption that the cooling function is a simple power-law of density and temperature. This choice ensures that the time evolution of the system is self-similar by requiring that the cooling time of the system at all times is a fixed fraction of the age of the universe. The solutions take into account many of the processes that are relevant to the assembly of the baryonic component of galaxies in a cosmological scenario: gravitational infall, energy dissipation through shocks, pressure gradients, radiative energy losses, and cooling flows. Analytic solutions such as the ones outlined here are invaluable to assess the reliability and diagnose the shortcomings of numerical techniques currently being used in cosmological simulations. The tests we present here show that SPH simulations reproduce the analytic solutions very well. No substantial deviations from the predicted central mass accretion rates or from the temperature, density, and velocity profiles are observed inside the shock radius when cooling is included. The region affected by the shock is, however, is much larger than predicted: effects from the shock are seen as far out as 2 or 3 times the nominal shock radius. Although this does not seem to affect adversely the post-shock behaviour of the gas under the simplifying conditions we adopt, it may have unwanted consequences in cases with more complex infall geometry (such as mergers) or that involve a cooling function with a more sensitive dependence on temperature and density than assumed here. We hope that the work reported here will encourage further efforts to test and improve the numerical treatment of the hydrodynamics of galaxy formation. \bigskip \noindent This work has been partially supported by the National Science and Engineering Research Council of Canada. RGB acknowledges the support of the PPARC rolling grant ``Extragalactic Astronomy and Cosmology at Durham'' and the use of STARLINK computing facilities. MGA acknowledges the hospitality of the Physics Department of Durham University.
\section{Introduction} The relatively bright X--ray source 1RXS\,J170849.0--400910\ was originally discovered in the {\em ROSAT}\ All Sky Survey (RASS, Voges et al. 1998), but it did not attract much attention until $\sim$11\,s periodic pulsations were found during an ASCA observation that was part of the Galactic Plane Survey Project (Sugizaki et al. 1997). Based on the spin period value and the soft X--ray spectrum Sugizaki et al. (1997) suggested that this optically unidentified source could be a new member of the small group of AXPs. AXPs are a group of X--ray pulsars (Mereghetti \& Stella 1995; van Paradijs et al. 1995) that share several peculiar properties: a narrow interval of pulse periods (6--12 s), very soft X--ray spectra (a steep power--law with photon index $\sim$ 3--4 often coupled to an additional blackbody--like component with kT$\sim$ 0.4--0.6 keV), lower X--ray luminosities ($\lower.5ex\hbox{\ltsima}$10$^{35}$ erg s$^{-1}$) than those of persistent High Mass X--ray Binaries (HMXBs), a narrow spatial distribution in the Galactic plane, absence of large flux variations on timescales from months to years, spin--down trends at relatively stable rates (2--100 $\times$ 10$^{-5}$ s yr$^{-1}$), and absence of massive companion stars (see Mereghetti et al. 1998 for a recent review). Whether AXPs are isolated neutron stars or members of binary systems remains to be determined; similarly it is unclear whether they are powered by accretion or by magnetic energy, as recently suggested by the analogy with Soft $\gamma$--ray Repeaters (SGRs; Kouveliotou et al. 1998, Mereghetti 1999). In this Letter we report on the data analysis of public {\em ROSAT}\ HRI observations of 1RXS\,J170849.0--400910\ and on a study of its possible optical counterparts. Our results strongly support the suggestion that 1RXS\,J170849.0--400910\ is a member of the AXP group. \section{ROSAT HRI Observation} The field of 1RXS\,J170849.0--400910\ was observed twice by the {\em ROSAT}\ High Resolution Imager (0.1--2.4 keV): on 1994 March 8 (ROR 900595) for a total exposure time of 1120\,s and on 1997 March 24--26 (ROR 180188) for a total exposure time of 11175\,s. The source was detected at a level of 0.13$\pm$0.01 and 0.130$\pm$0.003 counts s$^{-1}$ in the 1994 and 1997 observations, respectively (90\% uncertainties are used through out this letter). We analyzed the two HRI images using both a sliding cell and a wavelet transform--based detection algorithm (Lazzati et al. 1999; Campana et al. 1999). In the longer observation (1997) the source was detected at R.A.=17$^h$08$^m$47$^s$.24 and DEC=--40$^o$08\arcmin50\arcsec.7 (J2000), with a statistical error of only 0.15\arcsec\ (1$\sigma$). However, due to the the uncertainty in the satellite boresight, an overall error radius of $\sim$ 10\arcsec\ must be considered. The new position is located some 35\arcsec\ and 25\arcsec\ North of the ASCA position (30\arcsec\ radius; Sugizaki et al. 1997) and PSPC RASS position (9\arcsec\ uncertainty). The 1993 {\em ROSAT}\ HRI observation was analyzed in the same way and provided a slightly different (but consistent) position, R.A.=17$^h$08$^m$46$^s$.52 and DEC=--40$^o$08\arcmin52\arcsec.5 (J2000; $\sim$ 10\arcsec\ radius uncertainty; see also Fig.~3). The {\em ROSAT}\ HRI has no intrinsic spectral resolution, therefore in order to obtain the source flux from the observed count rate we assumed the spectral model derived with ASCA (Sugizaki et al. 1997), i.e. a power--law with photon index $\Gamma$=3.45 and N$_H$=1.8$\times$10$^{22}$ cm$^{-2}$. In this way we determined a flux at the Earth of 1.1$\times$10$^{-11}$ erg cm$^{-2}$ s$^{-1}$ in the 0.1--2.4 keV energy range. For a distance of 10 kpc, this corresponds to a source unabsorbed luminosity of $L_X \sim$ 1.2 $\times$ 10$^{36}$ erg s$^{-1}$ in the 0.8--10 keV range. The event list of 1RXS\,J170849.0--400910\ was extracted from a circle of $\sim$40\arcsec\ radius (corresponding to an encircled energy of $\sim$90\%) around the X--ray position of the 1997 observation. The photon arrival times were corrected to the barycentre of the solar system and a background subtracted light curve was accumulated in 1\,s bins. A power spectrum was calculated over the entire observation duration. A highly significant peak ($\sim$ 8$\sigma$) was found at a frequency of 0.09092712 Hz, corresponding to a period of 10.998\,s. To refine the period determination and reduce its uncertainty we adopted a phase fitting technique. The best pulse period was determined to be P=10.99802$\pm$0.00005 \,s. The shape of the modulation is somewhat asymmetric with a pulsed fraction of $\sim$38$\pm$4\% (Fig.\,1). {\rc We note that such pulse shape is similar to that reported by Sugizaki et al. in the 0.8--2 keV range (1997; see their Fig.\,3). } Owing to poor statistics the periodicity was not detected during the 1994 observation {\rc (assuming a 3$\sigma$ detection threshold)}. \begin{inlinefigure} \bigskip \centerline{\includegraphics[width=0.53\linewidth]{RXS1708pul_apje.ps}} \caption{The 1997 March 24--26 {\em ROSAT}\ HRI light curve folded at the best period P=10.99802\,s.} \bigskip \end{inlinefigure} \begin{inlinefigure} \bigskip \centerline{\includegraphics[width=0.64\linewidth]{RXS1708_hr_apje.ps}} \bigskip \caption{Color--Magnitude Diagram for the objects in the field of 1RXS\,J170849.0--400910. The Main Sequence lines for different values of the distance and reddening are also shown (700\,pc, A$_{\rm V}$=1.6; 3 kpc, A$_{\rm V}$=6.2). The straight line indicates the direction of the reddening and corresponds to a value of A$_{\rm V}$=2. The squares mark the objects within the error circles.} \end{inlinefigure} By comparing our period value with the one measured with ASCA on 1996 September 3 (Sugizaki et al. 1997) a period derivative \.P = (7$\pm$2)$\times$10$^{-4}$ s yr$^{-1}$ was inferred. \\ \section{Optical Imaging and Spectroscopy} Photometry for each stellar object in the images was derived by means of DAOPHOT\,II (Stetson 1987), and a Color--Magnitude Diagram (CMD) was then computed (see Fig.\,2) for all stars within a region of 4\arcmin$\times$5\arcmin\ around the X--ray position of 1RXS\,J170849.0--400910. For comparison Fig.\,2 also shows the main sequence for two representative distances: 700\,pc (with A$_{\rm V}$=1.6) and 3\,kpc (A$_{\rm V}$=6.2). Note that 1RXS\,J170849.0--400910\ is located in the Galactic plane (l$^{II}$=346.48, b$^{II}$=0.03) and that along this direction spiral arms are located at a distance of $\sim$\,1kpc, $\sim$ 3\,kpc and $\sim$4.5\,kpc (Taylor \& Cordes 1993). The R image of the field of 1RXS\,J170849.0--400910\ is shown in Fig.\,3. The new X--ray position uncertainty regions obtained with the two HRI observations are shown (10$\arcsec$ radius). Table\,1 gives the positions, R magnitudes and the colors of the objects within (or close to) the error circles. The spectrum of the brightest object within the error circle (candidate A, R = 13.4; V--R = 1.48) is shown in Fig.\,4. The object was classified as a B0V spectral--type star with an absorption A$_{\rm V}\sim$6.2 . We found no evidence for any emission--lines in its spectrum. \begin{inlinefigure} \bigskip \centerline{\includegraphics[width=0.83\linewidth]{RXS1708_R_apje.ps}} \bigskip \caption{R image of the field of 1RXS\,J170849.0--400910. The new X--ray position uncertainty regions (10\arcsec\ radius) obtained with the 1994 and 1997 HRI observations are also shown. North is top, east is left.} \end{inlinefigure} \section{Discussion} The great majority of accreting X--ray pulsars are found in two distinct classes of massive binaries: systems with OB type supergiant companions and systems with Be stars. The latter are often characterized by transient activity. We found that star A, the brightest candidate counterpart in the new error circles, is a main sequence B0 star, but the properties of its spectrum make an association with 1RXS\,J170849.0--400910\ very unlikely. The absence of emission--lines corresponding to the Balmer series, HeI and FeII strongly suggest that star\,A is not in a binary system with a compact object. Although the optical emission--lines in Be/neutron star systems may occasionally disappear during the quiescence state, such a behaviour has been so far observed only in transient sources which show also outbursts and pronounced spin--up/down rate episodes. The low value of the inferred spin--down period derivative over a 6 month interval and the constant X--ray flux value over a timescale of 3 years (see below) rule against this possibility. Furthermore, the soft X--ray spectrum of 1RXS\,J170849.0--400910\ is much softer than that observed in all X--ray pulsars with massive companions (i.e. a hard power law with photon index $\sim$1 that steepens only above the cut--off energy of $\sim$10--30 keV). \begin{figure*}[!tb] \psfig{figure=CandA_apje.ps,width=11cm,height=5cm} \caption{4300--6800 \AA\ low--resolution flux--calibrated spectrum of the reddened B0V spectral--type star A. I.S. stands for inter--stellar absorption.} \end{figure*} \begin{inlinetable} \begin{center} \caption{Positions, R magnitudes and colors of the stars within (or close to) the X--ray error circles of 1RXS\,J170849.0--400910\ } \begin{tabular}{ccccc} \\ \hline \hline Star & R.A. (J2000)$^a$ & DEC (J2000)$^a$ & R$^b$ & V--R$^b$ \\ & (hh mm ss) & ($^\circ$ $\arcmin$ $\arcsec$) & mag & mag \\ \hline A & 17 08 46.2& --40 08 48& 13.41 & 1.48\\ B & 17 08 47.8& --40 08 40& 16.46 & 0.85\\ C & 17 08 46.8& --40 08 40& 17.36 & 1.48\\ D & 17 08 47.1& --40 08 36& 18.24 & 1.50\\ E & 17 08 47.6& --40 08 46& 18.26 & 1.27\\ F & 17 08 46.1& --40 08 43& 18.36 & 1.38\\ G & 17 08 47.0& --40 08 38& 18.60 & 1.50\\ H & 17 08 46.9& --40 09 05& 18.74 & 1.25\\ I & 17 08 46.4& --40 08 53& 18.86 & 1.61\\ J & 17 08 46.8& --40 08 50& 18.93 & 1.36\\ K & 17 08 47.8& --40 08 52& 19.00 & 1.38\\ L & 17 08 47.4& --40 08 53& 19.78 & 1.58\\ \hline \end{tabular} \end{center} \begin{minipage}{0.87\linewidth} \textsuperscript{a}--- Uncertainty $\sim$1\arcsec. \\ \textsuperscript{b}--- Absolute uncertainty $\sim$0.2 mag. Relative uncertainty $\sim$0.02 mag. \end{minipage} \end{inlinetable} The second brightest candidate to be considered is star B. Its position in the CMD of Fig.~2 is consistent with a relatively closely K type star. All remaining candidates are fainter than R$\sim$17 implying X--ray to optical flux ratios higher than those typically observed in High Mass X--ray Binaries. This is also indicated by their position in the CMD shown in Fig.~2: the objects in the error circles are too blue and too faint to be compatible with early type stars. Note also that our images are deep enough to exclude very reddened and distant O and B stars. The absorption column density for 1RXS\,J170849.0--400910\ is $\sim 1.5\times10^{22}$\,cm$^{-2}$ (Sugizaki et al. 1997), corresponding to E(B--V) $\sim 2.6$ (Bohlin et al. 1978). Applying the standard interstellar extinction law (Fitzpatrick 1998) this translates to an absorption in the R band of A$_{\rm R}\sim 6$. A B0 star located at a distance of $5 \div 10$\,kpc would be observed with a R magnitude $\sim 15.5 \div 17$ and V--R $\sim 2$, i.e. it should have been easily detected in our optical images. Our optical data cannot exclude that 1RXS\,J170849.0--400910\ is part of a Low Mass X--ray Binary (LMXB). The faint stars in the error box are all compatible with relatively late spectral--type stars. For example star\,I is consistent with a A5--7V at a distance of 3\,kpc, while star\,H is consistent with a late K or an early M at a distance of 700\,pc. However, none of these objects has the characteristic blue color generally observed in the counterparts of luminous LMXBs, in which the optical emission is dominated by the accretion disk (van Paradijs 1998). The lack of a massive companion star and the inferred spin--down at a rate of $\sim$7.5$\times$10$^{-4}$ s yr$^{-1}$ strongly support the inclusion of 1RXS\,J170849.0--400910\ in the group of AXPs, as was previously suggested by Sugizaki et al. (1997) based only on the soft X--ray spectrum and 11\,s period. 1RXS\,J170849.0--400910\ was detected by the {\em ROSAT}\ HRI at the same flux level in both the 1994 and 1997 observations. The {\em ROSAT}\ flux extrapolated to the 0.8--10 keV range (2.3$\times$10$^{-11}$ erg cm$^{-2}$ s$^{-1}$; for the power--law best fit) is about a factor of 2 lower than that observed with ASCA (4.3$\times$10$^{-11}$ erg cm$^{-2}$ s$^{-1}$). A soft blackbody component, accounting for 20--50\% of the total flux has been seen in the spectrum of three well studied AXPs (4U\,0142+61; White et al. 1996; Israel et al. 1999, 2E\,2259+587; Corbet et al. 1995; Parmar et al. 1998, and 1E\,1048.1--5937; Corbet \& Mihara 1997; Oosterbroek et al. 1998). The presence of a similar component in 1RXS\,J170849.0--400910\ was found compatible with the ASCA data (although not formally required by the fit, Sugizaki et al. 1997). Assuming a double component spectral model for 1RXS\,J170849.0--400910, at least part of the flux difference between the ASCA and {\em ROSAT}\ observations might be ascribed to the more complex spectral slope than assumed. We conclude that the modest luminosity variations across observations spanning three years is another analogy of 1RXS\,J170849.0--400910\ with the AXPs class. {\rc 1RXS\,J170849.0--400910, like other AXPs, displays also a remarkable stability of the pulsed fraction and pulse shape.} One of the main problems in understanding the nature of AXPs is to establish whether they are isolated neutron stars or members of binary systems with very low mass companions. Both deep optical and infrared searches for counterparts and searches for Doppler orbital modulations in the pulse frequency can provide valuable information in this respect. In the best studied AXPs, the limits derived from such studies exclude most classes of companions stars, but helium stars with M$\leq 0.8 M_{\odot}$ and white dwarves (Mereghetti, Israel \& Stella 1998; Wilson et al. 1999). The possibility that 1RXS\,J170849.0--400910\ has a white dwarf companion star is compatible with our optical observations: the expected R magnitude at a distance of 700\,pc would be $\geq$ 21, well below the limiting magnitude of our images. Similar results apply to a helium main sequence star. We also note that although the absorption column inferred with the ASCA data would place 1RXS\,J170849.0--400910\ quite far from us (10 kpc), it is possible that part of the absorption be intrinsic. {\rc Also the possibility that 1RXS\,J170849.0--400910\ is a ``magnetar'', an isolated strongly magnetic neutron star, possibly related to Soft $\gamma$--ray Repeaters remains viable. More period measurements are clearly required to look for ``glitches'' as foreseen in the ``magnetar'' scenario.} In conclusion, based on the X--ray and optical findings presented here, we confirm that 1RXS\,J170849.0--400910\ is a new member of the AXPs group. \acknowledgments GLI and SC thanks V.F. Polcaro for his kind help and useful discussions. This work was partially supported through ASI grants. \vspace{3mm}
\section{Introduction} Solutions containing macromolecules are ubiquitous in the everyday life. From food colloids to the DNA, we are surrounded by these giant molecules which directly or indirectly govern every aspect of our lives. In many cases the macromolecules in solution posses a net charge. The electrostatic repulsion between the polyions is, often, essential to stabilization of colloidal suspensions. In the biological realm the electrostatics is responsible for the condensation of the DNA and formation of actin bundles, while various physiological mechanisms depend on the electrostatic interactions between the proteins and the microions. In spite of their ubiquity, our understanding of polyelectrolyte solutions is far from complete. The effort to fathom the role of electrostatics as it applies to the colloidal suspensions goes back over half a century to the classic works of Derjaguin and Landau~\cite{DL41} and of Verwey and Overbeek (DLVO).\cite{DLVO48} These in turn were based on the pioneering studies of Gouy~\cite{Go10} and Chapman~\cite{Ch13} of double layers in metal electrodes. Following these early contributions, a large effort has been devoted to solve the Poisson-Boltzmann (PB) equation in various geometries. The mean-field treatment, based on the solution of the PB equation, suggests that the interaction between two equally charged macroions in a suspension containing counterions is always repulsive.\cite{Is92,Sa94} In recent years, however, this dogma began to be questioned based on simulations,\cite{Gu84} analytical calculations~\cite{Kj84,Po89,Lo90,St90,Ro96,Pi98,Le99} and experiments,\cite{Is86,Isr86,Cr96,Ca96} which indicated that for small distances and large charge densities, two like-charged polyions might actually attract! The fundamental goal of this paper is to demonstrate that this attraction is linked to the correlations between the microions omitted in the mean-field theories, and to establish the conditions under which the attraction becomes possible. We shall consider the interaction between two infinite uniformly charged plates confining their own point-like counterions. The mean-field approximation for this system is obtained by solving the PB equation which, due to the planar symmetry, can be done analytically. Once the density profile is obtained, all the other thermodynamic quantities can be easily derived. Thus, it is not difficult to demonstrate that the pressure at the mean-field level, in units of energy, is simply the density of counterions at the mid-plane between the plates. Since this is always positive, no attraction is possible within the mean-field theory. Realization that the correlations between the counterions can strongly modify the mean-field predictions goes back a number of years. One of the first approaches proposed by Kjellander and Mar\v{c}elja~\cite{Kj84} was to include the correlations through the numerical solution of the Anisotropic Hypernetted Chain Equation (AHNC). These authors found that the force per unit area (pressure) can become negative in the presence of divalent counterions. Monte Carlo (MC) simulations performed by Guldbrand {\it et al.}~\cite{Gu84} also indicate that as the surface charge density is increased, the pressure decreases if the distance between the charged surfaces is sufficiently small. As in the case of the AHNC calculations, attraction was found only in the presence of divalent counterions. These authors, however, did not analyze the case of very high charge density and short distance between the plates. In addition, since in the above calculations it is difficult to separate the different physical contributions to the pressure, the mechanism that drives the attraction remains unclear. A different theoretical approach which attempted to shed some light on the mechanism of attraction was advanced by Stevens and Robbins.\cite{St90} These authors proposed a density-functional theory similar to the one often employed in studies of simple liquids. This approach introduces a grand-potential free energy, $\Omega \left[\rho(\mbox{\boldmath$r$})\right]$, which is a functional of the non-uniform density of counterions $\rho(\mbox{\boldmath$r$})$. The equilibrium properties of the system are obtained through the minimization of the total free energy. The practical problem with this method is that the exact form of the functional is not known. When the correlations between the microions are omitted, the minimization of the grand potential, $\Omega_{\rm PB}$, becomes trivial and leads to the usual PB equation.\cite{Sa94} In order to account for the correlations between the counterions, Stevens and Robbins~\cite{St90} appealed to the Local Density Approximation (LDA).\cite{Ta85,Cu85,De89} Within this approach an additional contribution, $f_{\rm LDA}$, is added to the mean-field expression, $\Omega_{\rm PB}$. The expression for $f_{\rm LDA}$ adopted by Stevens and Robbins was obtained through the extrapolation of the MC data for the {\it homogeneous}\/ One-Component Plasma (OCP),\cite{Br79} but with the homogeneous density replaced by an {\it inhomogeneous}\/ density profile. The minimization of the free-energy functional allowed them to determine the density profile, $\rho(\mbox{\boldmath$r$})$, and the pressure, $P_{\rm OCP}$. The LDA, however, is not without its own problems. The major drawback of this approach is that, for short distances and high charge densities, the LDA is unstable. The reason for the instability is due to the fact that as the density of counterions in the vicinity of the plates increases, the chemical potential decreases, what attracts more particles to the region. This, in turn, leads to an unphysical ``chain reaction'' where all the counterions condense onto the plates. Clearly, when the distance between the counterions becomes smaller than some threshold value, $s_{\rm corr}$, the LDA ceases to be a reliable approximation.\cite{St90,Gr91,St96} An improvement over the LDA is, the so called, Weighted Density Approximation (WDA).\cite{Ta85,Cu85,De89,Gr91} In this case, the excess free energy is taken to be a function of an {\it average}\/ density, $\rho_{w}(\mbox{\boldmath$r$})=\int {\rm d}^3\mbox{\boldmath$r$}'\,w(|\mbox{\boldmath$r$}-\mbox{\boldmath$r$}'|)\,\rho(\mbox{\boldmath$r$}')$, averaged over a region of radius $s=s_{\rm corr}$, where the interactions between the counterions are the strongest.\cite{Cu85,De89} The difficulty in the practical implementation of this scheme is the determination of a proper weight function. The simplest possible form for $w(|\mbox{\boldmath$r$}-\mbox{\boldmath$r$}'|)$, used by Stevens and Robbins,\cite{St90,St96} was to assume that this function has a long-range variation comparable to the wall separation.\cite{Rob} In this case, the weighted density $\rho_w(\mbox{\boldmath$r$})$ is approximated by the homogeneous density independent of $\mbox{\boldmath$r$}$. However, when the walls are not close, $L>s_{\rm corr}$, the weighted function is no longer uniform and the approximation adopted by Stevens and Robbins becomes unrealistic. A beautiful explanation of the attraction between like-charged plates has been recently advanced by Rouzina and Bloomfield.\cite{Ro96} These authors present a picture of attraction as arising from the ground-state configuration of the counterions. Clearly at zero temperature the counterions will recondense onto the surface of the plates forming two intercalating Wigner crystals. The authors advance a hypothesis that even at finite temperatures, relevant to the common experimental conditions, the attraction is still governed by the zero-temperature correlations. A somewhat different formulations based on field-theoretic methodology have also been proposed. In these approximations the attraction arises as a result of correlated fluctuations in the counterion charge densities.\cite{Po89,Pi98,Le99} Although providing a nice qualitative explanation of the origin of the attraction, these simple theories fail to yield a quantitative agreement with the simulations. In this paper we propose a different form of the weighted-density approach, which rectifies the problems of the earlier theories while still remaining numerically tractable. The excess free energy and the weight function, $w(|\mbox{\boldmath$r$}-\mbox{\boldmath$r$}'|)$, are derived from the Debye-H{\"u}ckel-Hole (DHH) theory of the OCP.\cite{No84} The density profile is determined by minimizing the free-energy density with respect to the {\it local}\/ density. Once the density profile is obtained, the free energy of the system is calculated by inserting it into the expression for the free-energy functional. Given the free energy, all the thermodynamic properties of the system can be easily calculated. A careful analysis of the behavior of the pressure as a function of the charge density and the distance between the plates allows us to explore the nature and the origin of the attraction. The remainder of the paper is organized as follows. The model and the PB approximation for the density-functional approach are described in Sec.~\ref{poisson}. The WDA is introduced and applied in Sec.~\ref{wda}. Our results and conclusions are summarized in Sec.~\ref{conclusion}. \section{The Poisson-Boltzmann Approach} \label{poisson} We consider two large, charged, thin surfaces each of area ${\cal A}$, separated by a distance $L$ (see Fig.~\ref{model}). The two plates with a negative surface charge density, $-\sigma$, confine positive point-like monovalent counterions with charge $e$. The overall charge neutrality of the system is guaranteed by the constraint \begin{equation} \label{1} \int_{-L/2}^{L/2}{\rm d} z\,\rho(z)=\frac{2\sigma}{e}, \end{equation} where $\rho(z)$ is the local number density of counterions and $z$ is the Cartesian coordinate perpendicular to the plates. The space between the plates is assumed to be a dielectric continuum of constant $\varepsilon$. In order to explore the thermodynamic properties of the system, we use a density-functional approach. The grand potential of the system is \begin{equation} \label{2} \Omega \left[\rho\right] \equiv {\cal F}[\rho]-\mu N\;, \end{equation} where $N$ is the total number of counterions, $\mu$ is their chemical potential and the functional $\cal F$ is derived from the free-energy density of the homogeneous system, with the uniform density of counterions, $\rho_{\rm c}=N/L{\cal A}$, replaced by the local density $\rho(z)$. For dilute systems, the ionic correlations can be neglected and the grand-potential functional (per unit area) becomes \end{multicols} \ruleleft \begin{equation} \label{4} \frac{\beta {\Omega}\left[\rho \right]}{{\cal A}}= \int_{-L/2}^{L/2}{\rm d} z\,\rho(z)\left\{\ln \left[\Lambda^3 \rho(z)\right]-1\right\} +\frac{\beta}{2} \int_{-L/2}^{L/2}{\rm d} z\,\phi(z)\left[e\rho(z)+q(z)\right] -\beta \mu \int_{-L/2}^{L/2}{\rm d} z\, \rho(z)\;, \end{equation} \ruleright \begin{multicols}{2} where the electrostatic potential, \begin{equation} \label{4b} \phi(\mbox{\boldmath$r$})=\int {\rm d}^3\mbox{\boldmath$r$}'\,\frac{e\rho (\mbox{\boldmath$r$}')+q(\mbox{\boldmath$r$}')}{\varepsilon |\mbox{\boldmath$r$}-\mbox{\boldmath$r$}'|}, \end{equation} due to the symmetry of the problem, depends only on the $z$ coordinate. $\Lambda$ is the de Broglie thermal wavelength of the counterions, $\beta=1/k_B T$ and $q(z)=-\sigma\left[\delta(z-L/2)+\delta(z+L/2)\right]$ is the surface charge density of the plates. The functional minimization of this expression, \begin{equation} \label{5} \frac{1}{\cal A}\frac{\delta\beta \Omega}{\delta\rho(z)}=0\;, \end{equation} produces the optimum density profile, \begin{equation} \label{6} \rho(z)={\rho}_0\,\exp \left[-\beta e\phi(z)\right]\;. \end{equation} \begin{figure}[ht] \vspace*{2.5cm} \begin{center} \leavevmode \epsfxsize=0.47\textwidth \epsfbox[25 20 580 450]{ model.ps"} \end{center} \vspace*{-2.5cm} \begin{minipage}{0.48\textwidth} \caption{Two infinite, negatively charged thin plates, with surface charge density $-\sigma$ separated by distance $L$. The counterions are confined to the region between the plates. The solvent is modeled as a uniform medium of dielectric constant $\varepsilon$.} \label{model} \end{minipage} \end{figure} The constant $\rho_0$ is determined from the overall charge-neutrality condition, Eq.~(\ref{1}), \begin{equation} \label{7} {\rho}_0\equiv \frac{2\sigma}{e\displaystyle \int_{-L/2}^{L/2}{\rm d} z\,\exp\left[-\beta e\phi(z)\right]}\;. \end{equation} The electrostatic potential is obtained by solving the Poisson equation, \begin{equation} \label{8} \frac{{\rm d}^2\phi(z)}{{\rm d} z^2}=-\frac{4\pi}{\varepsilon}\left[e\rho(z)+q(z)\right ], \end{equation} with the distribution of free ions given by Eq.~(\ref{6}). We find \begin{equation} \label{9} \phi(z)=\frac{1}{\beta e}\ln \left[\cos^2 \left(\frac{z-z_0}{\lambda} \right)\right]-\phi_0\;, \end{equation} where $\phi_0$ is the reference potential, which we will set to zero. Here $\lambda=1/\sqrt{2\pi\lambda_B\rho_0}$ and $\lambda_B= \beta e^2/\varepsilon$ is the Bjerrum length. Eq.~(\ref{8}) has to obey two boundary conditions, namely, \begin{eqnarray} \label{10} E(z=0)&=&0\;,\nonumber\\ E\left(z=\pm \frac{L}{2}\right)&=&\pm\frac{4\pi\sigma}{\varepsilon}\;. \end{eqnarray} From the first equation, the electric field vanishes at the mid-plane and, therefore, $z_0=0$. The second equation imposes the discontinuity of the electric field at both charged surfaces, leading to \begin{equation} \label{11} \frac{1}{\lambda}\tan \left(\frac{L}{2\lambda}\right)=\frac{2\pi\sigma\lambda_B }{e}\;. \end{equation} The potential at a point $z$ is, then, given by \begin{equation} \label{12} \phi(z)=\frac{1}{\beta e}\ln \left[\cos^2 \left(\frac{z}{\lambda}\right)\right] \;, \end{equation} with $\lambda$ the root of Eq.~(\ref{11}). The optimum density profile derived from this potential, \begin{equation} \label{13} \rho(z)=\frac{{\rho}_0}{\cos^2(z/\lambda)}\;, \end{equation} can now be substituted into the free-energy functional, allowing the calculation of the total free energy. The thermodynamic properties of the system can be determined from a suitable differentiation of the total free energy. For example, the force between the two plates is given by the minus derivative of the free energy with respect to the separation $L$ between the two surfaces. This differentiation leads to a particularly simple expression for the force per unit of area (or pressure), \begin{equation} \label{14} \beta P={\rho}_0\;. \end{equation} We note that although it might be tempting to attribute this simple result to the contact theorem, this is not the case, since the conditions under which this theorem holds are violated in the present geometry; Eq.~(\ref{14}) is purely a mean-field result. \section{The Weighted-Density Approximation} \label{wda} For dense systems, the correlations between the microions become relevant. For instance, if a counterion is present at the position $\mbox{\boldmath$r$}$, due to the electrostatic repulsion, the probability that another counterion is located in its vicinity is drastically reduced. The correlations in the positions of the counterions reduce the mean-field estimate of the electrostatic free energy. No exact method exists for calculating this excess contribution. The simplest approximation, the LDA, consists of adding to the Eq.~(\ref{4}) a {\it local}\/ functional, \begin{equation} \label{15} f_{\rm LDA}=\int_{-L/2}^{L/2} {\rm d} z\, \rho(z) f_{\rm corr}\left[\rho(z)\right]\;, \end{equation} where $f_{\rm corr}\left[\rho(z)\right]$ is the correlational free energy per particle. Within the LDA one normally uses the expression derived for the homogeneous system, in which the uniform density $\rho_{\rm c}=N/L{\cal A}$ is replaced by the local density profile $\rho(\mbox{\boldmath$r$})$. Unfortunately, as was mentioned above, the LDA is unstable when the one-particle density $\rho(\mbox{\boldmath$r$})$ is a rapidly varying function of the position. For example, for high surface charge densities, the minimization of the grand potential has no solution.\cite{Gr91} To circumvent this and related problems intrinsic to the LDA, Tarazona~\cite{Ta85} and Curtin and Ashcroft~\cite{Cu85} proposed a WDA, in which the free-energy density, $f_{\rm LDA}$, is replaced by \begin{equation} \label{16} f_{\rm WDA}=\int_{-L/2}^{L/2} {\rm d} z\, \rho(z) f_{\rm corr}\left[\rho_w(z)\right]\;. \end{equation} The fundamental difference between the LDA and the WDA is that the latter is assumed to depend not on the local density $\rho(\mbox{\boldmath$r$})$, but on some average density within the neighborhood of the point $\mbox{\boldmath$r$}$, \begin{equation} \label{17} \rho_{w}(\mbox{\boldmath$r$})=\int{\rm d}^3\mbox{\boldmath$r$}' w\left[|\mbox{\boldmath$r$}-\mbox{\boldmath$r$}'| ; \rho(\mbox{\boldmath$r$})\right]\, \rho(\mbox{\boldmath$r$}')\; . \end{equation} This provides a control mechanism which prevents an unphysical, singular, buildup of concentration at one point. The grand potential is obtained by adding the excess free energy per area, given by Eq.~(\ref{16}), to the Eq.~(\ref{4}), \begin{eqnarray} \label{18} \frac{\beta\Omega \left[\rho\right]}{{\cal A}}&=&\int_{-L/2}^{L/2}{\rm d} z\, \rho(z)\left\{\ln \left[\Lambda^3 \rho(z)\right]-1\right\}\nonumber\\ &&+\frac{\beta}{2}\int_{-L/2}^{L/2}{\rm d} z\,\phi(z)\left[e\rho(z)+q(z)\right]\nonumber\\ &&+\beta\int_{-L/2}^{L/2} {\rm d} z\, \rho(z) f_{\rm corr} \left[\rho_w(z)\right]\nonumber\\ &&-\beta\mu\int_{-L/2}^{L/2}{\rm d} z\,\rho(z)\;. \end{eqnarray} Minimization of this expression leads to the optimum particle number density, \begin{equation} \label{19} \rho(z)={\rho}_0 \exp \left[-\beta e\phi(z)-\beta \mu_{\rm ex}(z)\right]\;, \end{equation} where the excess chemical potential derived from $f_{\rm WDA}$, Eq.~(\ref{16}), is \begin{eqnarray} \label{20} \mu_{\rm ex}(z)&=&\frac{\delta f_{\rm WDA}}{\delta \rho(z)}\nonumber\\ &=&f_{\rm corr}\left[\rho_w(z)\right]+ \int_{-L/2}^{L/2}{\rm d} z'\,\rho(z') \,\frac{\delta f_{\rm corr}\left[\rho_w(z')\right]}{\delta \rho(z)}\;,\nonumber\\ \end{eqnarray} and the normalization coefficient is \begin{equation} \label{20b} \rho_0\equiv \frac{2\sigma}{e\displaystyle \int_{-L/2}^{L/2}{\rm d} z\,\exp\left[-\beta e\phi(z) - \beta \mu_{\rm ex}(z)\right]} \; . \end{equation} The electrostatic potential satisfies the Poisson equation, Eq.~(\ref{8}), with the charge density given by the Eq.~(\ref{19}). Integrating the Poisson equation over a rectangular shell of area ${\cal A}$ and width $z$, and appealing to the Gauss' theorem, an integro-differential equation for the electric field $E(z)$ can be obtained, \begin{equation} \label{21} {\cal E}({\bar z})=4\pi{\bar \sigma}\frac{\displaystyle \int_{0}^{\bar z}{\rm d}{\bar z}' \exp\left[-{\bar \mu}_{\rm ex}({\bar z}')+\int_{0}^{{\bar z}'}{\rm d}{\bar z}''{\cal E}({\bar z}'')\right]} {\displaystyle \int_{0}^{{\bar L}/2} {\rm d}{\bar z}' \exp\left[-{\bar \mu}_{\rm ex}({\bar z}') +\int_{0}^{{\bar z}'}{\rm d}{\bar z}'' {\cal E}({\bar z}'')\right]}\;, \end{equation} where ${\cal E} \equiv e\beta\lambda_B E$, $\bar{\sigma}\equiv \sigma \lambda_B^2/e$, ${\bar z}\equiv z/\lambda_B$, ${\bar L}\equiv L/\lambda_B$ and $\bar{\mu}_{\rm ex}\equiv \beta\mu_{\rm ex}$. The local density $\rho(z)$, which enters in the calculation of the excess chemical potential, Eq.~(\ref{20}), can be obtained from the derivative of the electric field, since $\nabla \cdot\mbox{\boldmath$E$}(\mbox{\boldmath$r$})=4 \pi e \rho(\mbox{\boldmath$r$})/\varepsilon$. The Eq.~(\ref{21}) explicitly fulfills the two boundary conditions: ${\cal E}(0)=0$ and ${\cal E}(\pm {\bar L}/2)=\pm 4\pi\bar{\sigma}$. The solution of this equation depends on the specific form of the excess free-energy density and the weight function $w\left(|\mbox{\boldmath$r$}-\mbox{\boldmath$r$}'|\right)$. For the homogeneous OCP the electrostatic free energy can be easily obtained using the DHH theory of Nordholm.\cite{No84} This is a simple linear theory based on the ideas of Debye and H{\"u}ckel. The electrostatic potential of the OCP is assumed to satisfy a linearized PB equation. As a correction for the linearization, Nordholm postulated the existence of an excluded-volume region of size $s_{\rm corr}$, from which all other ions are excluded. The size of this region is such that the electrostatic repulsion between two counterions is comparable to the thermal energy. Recent calculations using a generalized Debye-H\"uckel theory indicate that this exclusion region is responsible for the oscillations observed in the structure factor of the OCP at high couplings.\cite{Le299} Following Nordholm, we find \begin{equation} \label{24} s_{\rm corr}=\frac{1}{\kappa_D}\left(1+3\lambda_B \kappa_D\right)^{1/3}-\frac{1 }{\kappa_D}\; , \end{equation} where $\kappa_D=\sqrt{4\pi\lambda_B\rho_{\rm c}}$ is the inverse of the Debye length. The excess free energy per particle is calculated to be \end{multicols} \ruleleft \begin{equation} \label{25} \beta f_{\rm OCP}=\frac{1}{4}\left[1+\frac{2\pi}{3\sqrt{3}} +\ln \left(\frac{\omega^2+\omega+1}{3}\right)-\omega^2 -\frac{2}{\sqrt{3}}\tan^{-1}\left(\frac{2\omega+1}{\sqrt{3}}\right)\right]\;, \end{equation} \ruleright \begin{multicols}{2} where $\omega=(1+3\lambda_B\kappa_D)^{1/3}$. The correlational free energy per particle for the WDA, $f_{\rm corr}$, which appears in (\ref{20}), is obtained by replacing $\rho_{\rm c}$ by $\rho_w(z)$ in the expression (\ref{25}), that is, $f_{\rm corr}\left[\rho_w(z)\right]= f_{\rm OCP}\left[\rho_{\rm c}\to\rho_w(z)\right]$. To obtain the weighted function~\cite{Ta85,Cu85} we require that the second functional derivative of the free energy $\cal F$ in the limit of homogeneous densities, \begin{eqnarray} \label{22} \frac{\delta^2\beta{ \cal F}}{\delta \rho(\mbox{\boldmath$r$})\delta\rho(\mbox{\boldmath$r$}')} &=&\frac{\delta^3(\mbox{\boldmath$r$}-\mbox{\boldmath$r$}')}{\rho(\mbox{\boldmath$r$})} +w(|\mbox{\boldmath$r$}-\mbox{\boldmath$r$}'|)\, \frac{\delta\beta\mu_{\rm ex}(\mbox{\boldmath$r$})}{\delta\rho(\mbox{\boldmath$r$}')}\nonumber\\ &&+\frac{\lambda_B}{|\mbox{\boldmath$r$}-\mbox{\boldmath$r$}'|}\;, \end{eqnarray} produces the direct correlation function $C_2(\mbox{\boldmath$r$})$ of the homogeneous system, \begin{equation} \label{23} \frac{\delta^2\beta {\cal F}}{\delta \rho(\mbox{\boldmath$r$})\delta\rho(\mbox{\boldmath$r$}')} =\frac{\lambda_B}{|\mbox{\boldmath$r$}-\mbox{\boldmath$r$}'|}- C_2(\mbox{\boldmath$r$}-\mbox{\boldmath$r$}')\;. \end{equation} Following Groot,\cite{Gr91} we find that a reasonable approximation for the weight function is \begin{equation} \label{26} w(r)=w(|\mbox{\boldmath$r$}|)=\frac{3}{2\pi s_{\rm corr}^2}\left(\frac{1}{r}-\frac{1}{s_{\rm corr}}\right) \Theta(s_{\rm corr}-r)\;, \end{equation} where $\Theta(x)$ is the Heaviside step function. It is important to remember that the radius of the excluded-volume region, $s_{\rm corr}$, is now a function of the position, since the average density $\rho_{\rm c}$, which appears in Eq.~(\ref{24}), is replaced by $\rho(z)$, the local density of counterions, see Eq.~(\ref{17}). Taking advantage of the planar symmetry of the system, the expression for the weighted density can be written explicitly as a one-dimensional quadrature, \end{multicols} \ruleleft \begin{eqnarray} \label{27} \rho_w(z)&=& \frac{3}{s_{\rm corr}^2}\int_{-L/2}^{L/2}{\rm d} z'\;\rho(z') \int_0^\infty {\rm d} \varrho\,\varrho\left(\frac{1}{\sqrt{\varrho^2+(z-z')^2}}-\frac{1} {s_{\rm corr}}\right)\Theta\left(s_{\rm corr}-\sqrt{\varrho^2+(z-z')^2}\,\right)\nonumber\\ &=&\frac{3}{s_{\rm corr}^2}\int_{z_{<}}^{z_{>}}{\rm d} z'\;\rho(z') \int_0^{\sqrt{s_{\rm corr}^2-(z-z')^2}} {\rm d} \varrho\,\varrho\left(\frac{1}{\sqrt{\varrho^2+(z-z')^2}}-\frac1{s_{\rm corr}} \right)\nonumber\\ &=&\frac{3}{2s_{\rm corr}^3}\int_{z_{<}}^{z_{>}}{\rm d} z'\;\rho(|z'|) \left(s_{\rm corr}-|z-z'|\right)^2\;, \end{eqnarray} \ruleright \begin{multicols}{2} where $z_{<}\equiv \max(-L/2,z-s_{\rm corr})$, $z_{>}\equiv \min(L/2,z+s_{\rm corr})$ and $s_{\rm corr}$ is a function of $z$ through $\rho(z)$. \section{Results and Conclusions} \label{conclusion} Once $f_{\rm corr}$, $\mu_{\rm ex}(z)$ and $\rho_w(z)$ are defined, the electric field and, consequently, the optimum density profile can be determined from the numerical iteration of Eq.~(\ref{21}) until convergence is obtained. The Helmholtz free energy, $F$, associated with the optimum counterion distribution (\ref{19}), is determined by substituting it into the free-energy functional ${\cal F}$, \end{multicols} \ruleleft \begin{equation} \label{28} \frac{\beta F}{{\cal A}}=\frac{2\sigma}{e} \left[\ln \left(\Lambda^3{\rho}_0\right)-1\right] -\frac{\beta e}{2} \int_{-L/2}^{L/2} {\rm d} z\,\rho(z) \phi(z)- \beta\sigma\phi\left(\frac{L}{2}\right) -\int_{-L/2}^{L/2} {\rm d} z\,\rho(z) \left\{\beta\mu_{\rm ex}(z) -\beta f_{\rm corr}\left[\rho_w(z)\right]\right\} \;. \end{equation} \ruleright \begin{multicols}{2} Using this expression, the pressure, for different distances between the plates, $L$, and various charge densities, $\sigma$, can be easily obtained through numerical differentiation, \begin{equation} \label{29} P=-\frac{1}{\cal A}\frac{\partial {F}}{\partial L}\;, \end{equation} as shown in Fig.~\ref{wda_curvas}. When the charge density is below a threshold value, $\bar{\sigma}<\bar{\sigma}_{c}$, the dimensionless pressure, $\lambda_B^3 \beta P$, is always positive and a monotonically decreasing function of $\bar{L}$. Above the critical surface charge density the pressure exhibits a distinct minimum. In particular, we find that for sufficiently high surface charge densities the force between the two like-charged surfaces becomes negative, i.e. the two plates attract! \begin{figure}[ht] \begin{center} \leavevmode \epsfxsize=0.4\textwidth \epsfbox[25 20 580 450]{ wda_curvas.ps"} \end{center} \vspace*{.7cm} \begin{minipage}{0.48\textwidth} \caption{The reduced osmotic pressure as a function of the plate separation for various surface charge densities ${\bar \sigma}=\sigma \lambda_B^2/e$: 1 ($\triangle$), 5 ($\Box$) and 7 ($\bigcirc$). The solid line is the WDA and the dashed line is the PB approximation for the same values of $\bar \sigma$.} \label{wda_curvas} \end{minipage} \end{figure} In order to compare our results with other theories,\cite{Kj84,St90} we assumed that the dielectric medium between the plates is water at room temperature and, consequently, that the Bjerrum length is $\lambda_B=7.14\,${\AA}. The distance between the plates is fixed at 150 {\AA} and the inverse of the surface charge density, $\Sigma=e/\sigma$, is varied from 40 {\AA}$^2$ to 1000 {\AA}$^2$. Our results, illustrated in Fig.~\ref{stev}, show that for small surface charge densities the pressure increases almost linearly with the inverse charge density, $\Sigma$. In this case, since $P_{\rm corr} \ll P_{\rm PB}$, the pressure is dominated by the PB behavior. However, when the charge density becomes large, the slope of $P_{\rm WDA}$ increases due to the strong repulsion between the counterions. \begin{figure}[ht] \begin{center} \leavevmode \epsfxsize=0.4\textwidth \epsfbox[25 20 580 450]{ stev.ps"} \end{center} \vspace*{.4cm} \begin{minipage}{0.48\textwidth} \caption{Variation of $\beta P$ with $\Sigma=e/\sigma$ for $L=150$ {\AA}: from PB (dashed), WDA (solid) and AHNC ($\bullet$) from Ref.~[8].} \label{stev} \end{minipage} \end{figure} We also compare our calculations with the simulations of Guldbrand {\it et al.}\cite{Gu84} In this case, the distance between the plates is fixed at 21 {\AA} and the surface charge density is varied from 0.01 C/m$^2$ to 0.6 C/m$^2$, as shown in Fig.~\ref{guld}. \begin{figure}[ht] \begin{center} \leavevmode \epsfxsize=0.4\textwidth \epsfbox[25 20 580 450]{ guld.ps"} \end{center} \vspace*{.6cm} \begin{minipage}{0.48\textwidth} \caption{The osmotic pressure as a function of the surface charge density, when the distance between the plates is fixed at $L=21$ \AA, from PB (dashed) and WDA (solid). The circles ($\bullet$) are the data from Ref.~[7].} \label{guld} \end{minipage} \end{figure} \noindent When the density of counterions is small, $P_{\rm WDA}$ does not differ significantly from $P_{\rm PB}$. As the surface charge density is increased, the correlations among the counterions become relevant and $P_{\rm WDA}$ changes its slope and begins to decrease. Our results are in good agreement with the simulations, which also indicate that for a separation of 21 {\AA} the pressure exhibits a region where it decreases with increase in the surface charge density.\cite{Gu84} \acknowledgments We acknowledge the fruitful discussions with Marcelo Louzada-Cassou, Rudi Podgornik, Roland Kjellander and Stjepan Mar\v{c}elja. One of us, Marcia Barbosa, is particularly grateful for the useful discussion with Mark O. Robbins. This work was supported in part by CNPq --- Conselho Nacional de Desenvolvimento Cient{\'\i}fico e Tecnol{\'o}gico and FINEP --- Financiadora de Estudos e Projetos, Brazil. This research was also supported by the National Science Foundation under Grant No. PHY94-07194.
\section{Introduction} Over the past decade or so, many methods have been proposed for obtaining the three interior angles of the unitarity triangle, $\alpha$, $\beta$ and $\gamma$. In the near future these CP phases will be measured in a variety of experiments at $B$-factories, HERA-B, and hadron colliders. As always, the hope is that these measurements will reveal the presence of physics beyond the standard model (SM). The CP angles are typically extracted from CP-violating rate asymmetries in $B$ decays \cite{BCPasym}. New physics, if present, will affect these asymmetries principally through new contributions to loop-level processes. Most asymmetries involve the tree-level decays of neutral mesons ($B_d^0$ or $B_s^0$), and the new physics can enter into $B^0$-${\overline{B^0}}$ mixing \cite{NPBmixing}. However, it is occasionally the case that penguin contributions are involved, and the new physics can enter here as well \cite{GroWorah}. The canonical decay modes for measuring $\alpha$ and $\beta$ are $B_d^0(t) \to \pi^+\pi^-$ and $B_d^0(t) \to J/\Psi K_{\sss S}$, respectively. There have been many proposals for measuring the angle $\gamma$. One of these, accessible at asymmetric $e^+e^-$ $B$-factories, involves the CP asymmetry in $B^\pm \to D K^\pm$ \cite{BDK}. Another method, which is more appropriate to hadron colliders, uses $B_s^0(t) \to D_s^\pm K^\mp$ \cite{BsDsK}. How will we know whether or not new physics is present? One obvious way is if the three angles do not add up to $180^\circ$. Unfortunately, if the CP-angles are obtained in the conventional ways described above, $B$-factories can never find $\alpha+\beta+\gamma \ne \pi$. The reason is as follows: if there is new physics in $B_d^0$-${\overline{B_d^0}}$ mixing, the CP asymmetries in $B_d^0(t) \to \pi^+\pi^-$ and $B_d^0(t) \to J/\Psi K_{\sss S}$ will both be affected, but in opposite ways: instead of measuring $\alpha$ and $\beta$, the true (SM) CP-angles, one will extract $\tilde{\alpha } = \alpha - \theta_{\scriptscriptstyle NP}$ and $\tilde{\beta }= \beta + \theta_{\scriptscriptstyle NP}$, where $\theta_{\scriptscriptstyle NP}$ is the new-physics phase \cite{NirSilv}. On the other hand, since the measurement of $\gamma$ does not involve $B_d^0$ decays, it will be unaffected by new physics. Thus, even in the presence of new physics, one will still find $\tilde{\alpha} + \tilde{\beta} + \gamma = \pi$. $B$-factories must therefore find other ways of testing for the presence of new physics \cite{newphysics}. The most common method is simply to compare the unitarity triangle constructed from measurements of the angles with that constructed from independent measurements of the sides. If there is a discrepancy, one can then deduce that new physics is present, probably in $B_d^0$-${\overline{B_d^0}}$ mixing. The problem here is that there are large theoretical errors in obtaining the length of the sides of the unitarity triangle from the experimental data. Because of this, the presently-allowed region for the unitarity triangle is still rather large \cite{AliLondon}. Thus, if $\theta_{\scriptscriptstyle NP}$ is relatively small, one may still find agreement in the unitarity triangles constructed from the angles and sides. Furthermore, even if there is a discrepancy, one isn't sure whether new physics really is present --- it may simply be that the errors on the theoretical input quantities have been underestimated. The point here is that we would really like to find a method of directly probing new physics in $B_d^0$-${\overline{B_d^0}}$ mixing. Note that there are a variety of ways of testing for new physics in $B_s^0$-${\overline{B_s^0}}$ mixing, or, more precisely, in the $b\to s$ flavour-changing neutral current (FCNC). For example, above we mentioned two methods for obtaining $\gamma$: via $B^\pm \to D K^\pm$ and $B_s^0(t) \to D_s^\pm K^\mp$. If there is a discrepancy in the value of $\gamma$ obtained from these two decays, this would be a direct indication of new physics in $B_s^0$-${\overline{B_s^0}}$ mixing. A second example, somewhat different, is to measure $\beta$ using the decay $B_d^0(t) \to \phi K_{\sss S}$ \cite{LonSoni}. This is a pure $b\to s$ penguin decay. Thus, if the values of $\beta$ extracted via $B_d^0(t) \to J/\Psi K_{\sss S}$ and $B_d^0(t) \to \phi K_{\sss S}$ were to disagree, this would indicate the presence of new physics in the $b\to s$ penguin, i.e.\ in the $b\to s$ FCNC. (There might also be new physics in $B_d^0$-${\overline{B_d^0}}$ mixing, but the effect would be the same for the two decays.) Since new physics which affects $B_s^0$-${\overline{B_s^0}}$ mixing is also likely to affect the $b\to s$ penguin, in some sense this is a way of probing new physics in $B_s^0$-${\overline{B_s^0}}$ mixing without actually using $B_s^0$ mesons. Finally, a third example involves the CP asymmetry in $B_s^0(t) \to J/\Psi \phi$. To a good approximation, this asymmetry vanishes in the SM, so that a nonzero value would be clear evidence of new physics, specifically in $B_s^0$-${\overline{B_s^0}}$ mixing. Although there are many ways of getting at new physics in the $b\to s$ FCNC, to date no method has been suggested which directly tests for new physics in the $b\to d$ FCNC\footnote{In Ref.~\cite{Dalitz} it was claimed that the study of the Dalitz plot of $B_d^0(t)\to \pi^+\pi^-\pi^0$ decays allows one to cleanly perform such tests. However, it has since been shown that this particular point is in error, see Ref.~\cite{LSS}.}. In this Letter, we propose a method for doing just this. Essentially, the technique compares the weak phase of $B_d^0$-${\overline{B_d^0}}$ mixing with that of the $t$-quark contribution to the $b\to d$ penguin. In the SM, these phases are the same, since they involve the same Cabibbo-Kobayashi-Maskawa (CKM) \cite{CKM} matrix elements $V_{tb}^* V_{td}$. However, if there is new physics in the $b\to d$ FCNC, there may be a discrepancy. The method involves the decay $B_s^0 \to \phiK_{\sss S}$, along with its quark-level flavour-$SU(3)$ counterpart, $B_d^0 \to K^0 {\overline{K^0}}$. Our test of new physics in the $b\to d$ FCNC is not entirely clean -- it involves some theoretical input. However, the assumption we make is reasonably well-motivated, and so this may provide a first direct probe for new physics in the $b\to d$ FCNC. We also discuss a variation of this method involving the decays $B_s^0 \to J/\Psi K_{\sss S}$ and $B_d^0 \to J/\Psi \pi^0$. \section{$B_s^0(t)\to\phiK_{\sss S}$ and $B_d^0(t) \to K^0 {\overline{K^0}}$} When discussing the weak phases probed in various CP asymmetries, it is convenient to use approximate Wolfenstein parametrization of the CKM matrix \cite{Wolfenstein}, in which only $V_{td}$ and $V_{ub}$ have significant non-zero phases. The CKM phases in the unitarity triangle are then $\beta = {\rm Arg}(V_{td}^*)$ and $\gamma = {\rm Arg}(V_{ub}^*)$, with $\alpha$ defined to be $\pi - \beta - \gamma$. We will use this parametrization throughout the paper. We begin by considering the decay $B_s^0\to\phiK_{\sss S}$. The CP asymmetry in $B_s^0(t)\to\phiK_{\sss S}$ measures the relative phases of the two amplitudes ($B_s^0\to\phiK_{\sss S}$) and ($B_s^0\to{\overline{B_s^0}}$)(${\overline{B_s^0}}\to\phiK_{\sss S}$). To begin with, let us assume that there is no new physics. $B_s^0\to\phiK_{\sss S}$ is a pure penguin decay, which at the quark level takes the form ${\bar b} \to {\bar d} s {\bar s}$. Suppose first that this decay is dominated by an internal $t$-quark, in which case the CKM matrix-element combination involved is $V_{tb}^* V_{td}$. Since $B_s^0$-${\overline{B_s^0}}$ mixing involves $(V_{tb}^* V_{ts})^2$, the CP asymmetry probes \begin{equation} {\rm Arg} \left[ { V_{tb}^* V_{td} \over (V_{tb}^* V_{ts})^2 V_{tb} V_{td}^* } \right] = -2 \beta ~. \end{equation} Thus, within the SM, if the $t$-quark dominates the $b\to d$ penguin, one expects the CP asymmetry in $B_d^0(t) \to J/\Psi K_{\sss S}$ to be equal to that in $B_s^0(t) \to \phiK_{\sss S}$ \cite{LonPeccei}. If there is new physics, the CP asymmetry in $B_s^0(t)\to\phiK_{\sss S}$ can be affected in two ways: there may be new contributions to $B_s^0$-${\overline{B_s^0}}$ mixing (the $b\to s$ FCNC) and/or to the penguin decay $B_s^0\to\phiK_{\sss S}$ (the $b\to d$ FCNC). As discussed previously, new physics in $B_s^0$-${\overline{B_s^0}}$ mixing can be discovered independently, for example by comparing the CP asymmetries in $B^\pm \to D K^\pm$ and $B_s^0(t) \to D_s^\pm K^\mp$. If, after taking this into account, there is still a discrepancy in the value of $\beta$ as extracted from the CP asymmetries in $B_d^0(t) \to J/\Psi K_{\sss S}$ and $B_s^0(t) \to\phiK_{\sss S}$, this will indicate the presence of new physics in the $b\to d$ FCNC. Note also that one can perform a similar analysis with the related decay $B_d^0 \to K^0 {\overline{K^0}}$ \cite{LonPeccei}. (At the quark level, the decays are identical, save for the flavour of the spectator quark; at the meson level there is a difference since here there are two pseudoscalars in the final state, while the $B_s^0$ decay has a vector and a pseudoscalar.) Under the same assumptions as above, the CP asymmetry in $B_d^0(t) \to K^0 {\overline{K^0}}$ measures \begin{equation} {\rm Arg} \left[ { V_{tb}^* V_{td} \over (V_{tb}^* V_{td})^2 V_{tb} V_{td}^* } \right] = 0 ~. \end{equation} Thus, a non-zero CP asymmetry in this mode would indicate the presence of new physics in the $b\to d$ FCNC. However, there is a problem with the above analysis. Theoretical estimates suggest that the $b\to d$ penguin is {\it not} dominated by an internal $t$-quark. On the contrary, the $u$- and $c$-quark contributions can be substantial, perhaps even as large as 20\%--50\% of the $t$-quark contribution \cite{ucquark}. If this is the case, then the CP asymmetry in $B_s^0(t)\to\phiK_{\sss S}$ no longer cleanly probes the angle $\beta$. Instead, there is now ``penguin pollution'' from the $u$- and $c$-quark contributions to the $b\to d$ penguin, and the result depends on (unknown) hadronic quantities such as the strong phases and the relative sizes of the various penguin contributions. Thus, if one wants to detect new physics in the $b\to d$ FCNC, it is necessary to deal with this penguin pollution. As we show below, this can be done by combining information from both $B_s^0(t)\to\phiK_{\sss S}$ and $B_d^0 \to K^0 {\overline{K^0}}$. The $B_s$ and $B_d$ systems differ in that the width difference between the light and heavy $B_s$ eigenstates may be measurable, which is not the case for the $B_d$ system. In the presence of a non-negligible width difference, the expressions for the time-dependent decays of $B_s$ mesons are \cite{Dunietz} \begin{eqnarray} \label{Bsrates} \Gamma(B_s^0(t) \to f) & = & |A_f|^2 |g_+(t)|^2 + |{\bar A}_f|^2 |g_-(t)|^2 \nonumber \\ & ~ & \quad + 2 \left[ {\rm Re} (A_f^* {\bar A}_f) {\rm Re} (g_-(t) g_+^*(t)) - {\rm Im} (A_f^* {\bar A}_f) {\rm Im} (g_-(t) g_+^*(t)) \right] ~, \nonumber \\ \Gamma({\overline{B_s^0}}(t) \to f) & = & |{\bar A}_f|^2 |g_+(t)|^2 + |A_f|^2 |g_-(t)|^2 \nonumber \\ & ~ & \quad + 2 \left[ {\rm Re} (A_f^* {\bar A}_f) {\rm Re} (g_-(t) g_+^*(t)) + {\rm Im} (A_f^* {\bar A}_f) {\rm Im} (g_-(t) g_+^*(t)) \right] ~, \end{eqnarray} with \begin{eqnarray} |g_+(t)|^2 + |g_-(t)|^2 & = & {1\over 2} \left( e^{-\Gamma_{\scriptscriptstyle L} t} + e^{-\Gamma_{\scriptscriptstyle H} t} \right)~, \nonumber \\ |g_+(t)|^2 - |g_-(t)|^2 & = & e^{-\Gamma t} \cos\Delta m t ~, \nonumber \\ {\rm Re} [g_-(t) g_+^*(t)] & = & {1\over 4} \left( e^{-\Gamma_{\scriptscriptstyle L} t} - e^{-\Gamma_{\scriptscriptstyle H} t} \right) ~, \nonumber \\ {\rm Im} [g_-(t) g_+^*(t)] & = & {1\over 2} e^{-\Gamma t} \sin\Delta m t ~. \end{eqnarray} In the above, $\Gamma_{\scriptscriptstyle L}$ and $\Gamma_{\scriptscriptstyle H}$ are the widths of the light and heavy $B$-states, respectively, and $ \Gamma \equiv (\Gamma_{\scriptscriptstyle L} + \Gamma_{\scriptscriptstyle H}) / 2$. (Note also that we have assumed that the weak phase in $B_s^0$-${\overline{B_s^0}}$ mixing is zero, which holds within the SM.) Thus, the time-dependent measurements of $B_s$ decay rates allow one to obtain the following four functions of the decay amplitudes: \begin{equation} |A_f|^2 + |\bar{A}_f|^2 ~~,~~ |A_f|^2 - |\bar{A}_f|^2 ~~,~~ {\rm Re}\left( A_f^* \bar{A}_f \right) ~~,~~ {\rm Im}\left( A_f^* \bar{A}_f \right). \label{Aquantities} \end{equation} If the width difference in the $B_s$ system turns out to be small, then ${\rm Re} [g_-(t) g_+^*(t)] \simeq 0$, which appears to imply that one cannot obtain the quantity ${\rm Re}\left( A_f^* \bar{A}_f \right)$. However, this is not true. In fact, the above four functions are not independent, due to the equality \begin{equation} |A_f|^2 |{\bar A}_f|^2 = [{\rm Re} (A_f^* {\bar A}_f)]^2 + [{\rm Im} (A_f^* {\bar A}_f)]^2 ~. \end{equation} Thus, even if the width difference in the $B_s$ system is not measurable, we can still obtain each of the four quantities in Eq.~(\ref{Aquantities}), except that the {\it sign} of ${\rm Re}\left( A_f^* \bar{A}_f \right)$ is undetermined. The lack of knowledge of this sign simply leads to additional possible solutions (discrete ambiguities), one for each sign of ${\rm Re}\left( A_f^* \bar{A}_f \right)$. A measurable width difference allows the determination of this sign, thereby reducing discrete ambiguities. In what follows, we assume that the width difference is measurable, i.e.\ that the sign of ${\rm Re}\left( A_f^* \bar{A}_f \right)$ is known. If this sign cannot be determined, the method is still valid, but extra solutions are possible. Now, consider again the process $B_s^0\to\phiK_{\sss S}$. At the level of quark diagrams, there are several contributions to this decay: (i) the ordinary gluonic penguin, ${\tilde P}$, (ii) the Zweig-suppressed gluonic decay ${\tilde P}_1$, in which the gluon essentially hadronizes into the $\phi$, (iii) the electroweak penguin ${\tilde P}_{\scriptscriptstyle EW}$, and (iv) the colour-suppressed electroweak penguin ${\tilde P}_{\scriptscriptstyle EW}^{\scriptscriptstyle C}$. (The tildes on the amplitudes indicate a $B_s^0$ decay.) We therefore write the amplitude schematically as \begin{eqnarray} A_s^{\phi} &\equiv & A\left(B_s\rightarrow \phiK_{\sss S}\right) = {1\over \sqrt{2}} \left( {\tilde P} + {\tilde P}_1 + {\tilde P}_{\scriptscriptstyle EW} + {\tilde P}_{\scriptscriptstyle EW}^{\scriptscriptstyle C} \right) ~. \label{phiksamp} \end{eqnarray} (The factor $1/\sqrt{2}$ is included due to the presence of the $K_{\sss S}$.) Of these, the gluonic penguins receive contributions from internal $u$, $c$ and $t$-quarks, while the electroweak penguins are $t$-quark dominated. Any contribution to the $b\to d$ penguin can be written generically as \begin{eqnarray} P = \sum_{q=u,c,t} V_{qb}^* V_{qd} P_q & = & V_{cb}^* V_{cd} (P_c - P_u) + V_{tb}^* V_{td} (P_t - P_u) \nonumber \\ & \equiv & {\cal P}_{cu} e^{i\delta_c} + {\cal P}_{tu} e^{i\delta_t}e^{-i\beta} ~. \end{eqnarray} In the first line we have used the unitarity of the CKM matrix to eliminate the $u$-quark contribution, and in the second we have explicitly separated out the weak and strong phases and absorbed the magnitudes $|V_{cb}^* V_{cd}|$ and $|V_{tb}^* V_{td}|$ into the definitions of ${\cal P}_{cu}$ and ${\cal P}_{tu}$, respectively. Applying this to $B_s^0\to\phiK_{\sss S}$, Eq.~(\ref{phiksamp}) becomes \begin{equation} A_s^\phi = {1\over\sqrt{2}} \left( \tilde{\cal P}_{cu} e^{i{\tilde\delta}_c} + \tilde{\cal P}_{tu} e^{i{\tilde\delta}_t}e^{-i\beta } \right), \end{equation} where $\tilde{\cal P}_{cu}$ and $\tilde{\cal P}_{tu}$ are real and taken to be positive, and \begin{eqnarray} \tilde{\cal P}_{cu} e^{i{\tilde\delta}_c} & = & {\tilde P}_c - {\tilde P}_u + {\tilde P}_{1,c} - {\tilde P}_{1,u} ~, \nonumber \\ \tilde{\cal P}_{tu} e^{i{\tilde\delta}_t} e^{-i\beta } & = & {\tilde P}_t - {\tilde P}_u + {\tilde P}_{1,t} - {\tilde P}_{1,u} + {\tilde P}_{\scriptscriptstyle EW} + {\tilde P}_{\scriptscriptstyle EW}^{\scriptscriptstyle C} ~. \label{Ptildedefs} \end{eqnarray} With this expression for the $B_s^0\to\phiK_{\sss S}$ amplitude, the measurements of the quantities in Eq.~(\ref{Aquantities}) give \begin{eqnarray} \tilde{X} &\equiv & \frac{1}{2} \left( \left| A_s^{\phi}\right|^2 + \left| \bar{A}^{\phi}_s \right|^2 \right) = {1\over2} \left( \tilde{\cal P}_{cu}^2 +\tilde{\cal P}_{tu}^2 + 2 \tilde{\cal P}_{cu} \tilde{\cal P}_{tu} \cos{\tilde\Delta} \cos\beta \right) ~, \label{XP} \\ \tilde{Y} &\equiv & \frac{1}{2} \left( \left| A_s^{\phi}\right|^2 - \left| \bar{A}^{\phi}_s \right|^2 \right) = {1\over2} \left( - 2 \tilde{\cal P}_{cu} \tilde{\cal P}_{tu} \sin{\tilde\Delta} \sin\beta \right) ~, \label{YP} \\ \tilde{Z}_R &\equiv & Re\left( A_s^{\phi*} \bar{A}^{\phi}_{s } \right) = {1\over2} \left( - \tilde{\cal P}_{cu}^2 - \tilde{\cal P}_{tu}^2 \cos2\beta - 2 \tilde{\cal P}_{cu} \tilde{\cal P}_{tu} \cos{\tilde\Delta}\cos\beta \right) ~, \label{ZPR} \\ \tilde{Z}_I &\equiv & Im\left( A_s^{\phi*} \bar{A}^{\phi}_{s } \right) = {1\over2} \left( - \tilde{\cal P}_{tu}^2 \sin2\beta - 2 \tilde{\cal P}_{cu} \tilde{\cal P}_{tu} \cos{\tilde\Delta}\sin\beta \right) ~, \label{ZPI} \end{eqnarray} where ${\tilde\Delta}\equiv {\tilde\delta}_c - {\tilde\delta}_t$. (Note: in the above we have assumed that there is no new physics in $B_s^0$-${\overline{B_s^0}}$ mixing. As mentioned previously, the presence of such new physics can be independently determined. We will discuss the case of new physics in $B_s^0$-${\overline{B_s^0}}$ mixing further on.) Examining the above equations, we note that there are four unknown parameters ($\tilde{\cal P}_{cu}$, $\tilde{\cal P}_{tu}$, ${\tilde\Delta}$ and $\beta$), but only three independent measurements. Thus, we cannot solve for the unknowns. In particular, we see that we cannot obtain $\beta$ in the presence of penguin pollution. However, progress can be made if we also consider the decay $B_d^0 \to K^0 {\overline{K^0}}$, which is similar to $B_s^0\to\phiK_{\sss S}$. (At the quark level, they differ only in the flavour of the spectator quark.) For this decay, contributions come only from the ordinary gluonic penguin $P$ and the colour-suppressed electroweak penguin $P_{\scriptscriptstyle EW}^{\scriptscriptstyle C}$ (amplitudes without tildes indicate $B_d^0$ decays): \begin{equation} A_d^{\scriptscriptstyle KK} = A\left( B_d^0 \to K^0 {\overline{K^0}}\right) = P + P_{\scriptscriptstyle EW}^{\scriptscriptstyle C} ~. \end{equation} Analogous to the $B_s^0\to\phiK_{\sss S}$ amplitude, we can write \begin{equation} A_d^{\scriptscriptstyle KK} = {\cal P}_{cu} e^{i\delta_c} + {\cal P}_{tu} e^{i\delta_t}e^{-i\beta }, \end{equation} where \begin{eqnarray} {\cal P}_{cu} e^{i\delta_c} & = & P_c - P_u ~, \nonumber \\ {\cal P}_{tu} e^{i\delta_t} e^{-i\beta} & = & P_t - P_u + P_{\scriptscriptstyle EW}^{\scriptscriptstyle C} ~. \label{Pdefs} \end{eqnarray} In the $B_d^0$ system, the width difference is negligible, so that only three quantities can be obtained from time-dependent measurements: \begin{eqnarray} {X} &\equiv & \frac{1}{2} \left( \left| A_d^{\scriptscriptstyle KK}\right|^2 + \left| \bar{A}^{\scriptscriptstyle KK}_d \right|^2 \right) = {\cal P}_{cu}^{2} +{\cal P}_{tu}^{2} + 2 {\cal P}_{cu} {\cal P}_{tu} \cos\Delta \cos\beta ~, \label{XXP} \\ {Y} &\equiv & \frac{1}{2} \left( \left| A_d^{\scriptscriptstyle KK}\right|^2 - \left| \bar{A}^{\scriptscriptstyle KK}_d \right|^2 \right) = - 2 {\cal P}_{cu} {\cal P}_{tu} \sin \Delta \sin\beta ~, \label{YYP} \\ {Z} &\equiv & Im\left( e^{-2i \tilde{\beta }} A_d^{\scriptscriptstyle KK*} \bar{A}^{\scriptscriptstyle KK}_{d } \right) \nonumber \\ &~& \qquad\qquad = - {\cal P}_{cu}^{2} \sin 2\tilde{\beta } - {\cal P}_{tu}^{2} \sin(2 \tilde{\beta } - 2 \beta ) - 2 {\cal P}_{cu} {\cal P}_{tu} \cos\Delta \sin(2 \tilde{ \beta } - \beta ) ~, \label{ZZZP} \end{eqnarray} where $\Delta \equiv \delta_c - \delta_t$. Since we are allowing for the possibility of new physics in the $b\to d$ FCNC, we have explicitly denoted the weak phase of $B_d^0$-${\overline{B_d^0}}$ mixing as $\tilde\beta$, which may be different from the weak phase of the $b\to d$ penguin $\beta$. It is reasonable to assume that the mixing phase $\tilde\beta$ will be measured independently via $B_d^0(t) \to J/\Psi K_{\sss S}$. Even so, we are still left with three equations in four unknowns (${\cal P}_{cu}$, ${\cal P}_{tu}$, $\Delta$ and $\beta$), so once again we cannot solve for $\beta$. However, we can reduce the number of independent parameters in the $B_s^0(t)\to\phiK_{\sss S}$ and $B_d^0(t) \to K^0 {\overline{K^0}}$ measurements by making an assumption. Specifically, we assume that $r = \tilde{r}$, where $r \equiv {\cal P}_{cu}/{\cal P}_{tu}$ and $\tilde{r} \equiv \tilde{\cal P}_{cu}/\tilde{\cal P}_{tu}$. How good is this assumption? {}From Eqs.~(\ref{Ptildedefs}) and (\ref{Pdefs}) we have \begin{eqnarray} r & = & \left\vert { P_c - P_u \over P_t - P_u + P_{\scriptscriptstyle EW}^{\scriptscriptstyle C}} \right\vert \nonumber\\ {\tilde r} & = & \left\vert { {\tilde P}_c - {\tilde P}_u + {\tilde P}_{1,c} - {\tilde P}_{1,u} \over {\tilde P}_t - {\tilde P}_u + {\tilde P}_{1,t} - {\tilde P}_{1,u} + {\tilde P}_{\scriptscriptstyle EW} + {\tilde P}_{\scriptscriptstyle EW}^{\scriptscriptstyle C} } \right\vert ~. \end{eqnarray} Now, at the quark level the only difference between the $P$ and the ${\tilde P}$ amplitudes is the flavour of the spectator quark. Since this flavour should not have a significant effect on the size of the amplitude, for a given type of penguin contribution we can take $|P_i| \simeq |{\tilde P}_i|$. Furthermore, we can estimate the relative sizes of the various types of penguin contribution: $|{\tilde P}_{\scriptscriptstyle EW} / {\tilde P}_t| \simeq |{\tilde P}_{1,q} / {\tilde P}_q| \simeq |{\tilde P}_{\scriptscriptstyle EW}^{\scriptscriptstyle C} / {\tilde P}_{\scriptscriptstyle EW}| \simeq {\bar\lambda}$, where ${\bar\lambda} \sim 20\%$. Thus, we find that \begin{equation} r \simeq {\tilde r} = \left\vert { P_c - P_u \over P_t - P_u} \right\vert \end{equation} and \begin{equation} {r - {\tilde r} \over r} = O({\bar\lambda}) ~. \end{equation} Taking $r = \tilde{r}$ is therefore a reasonable assumption. With the assumption that $ r = \tilde{r}$, the measurements take the form \begin{eqnarray} \tilde{X} &=& {1\over 2} \, \tilde{\cal P}_{tu}^2 [ 1 + 2 r \cos\tilde{\Delta} \cos\beta + r^2 ] ~, \label{XBSPK} \\ \tilde{Y} &=& {1\over 2} \, \tilde{\cal P}_{tu}^2 [ - 2 r \sin\tilde{\Delta} \sin\beta ] ~, \label{YBSPK} \\ \tilde{Z}_R &=& {1\over 2} \, \tilde{\cal P}_{tu}^2 [ - \cos2\beta - 2 r \cos\tilde{\Delta}\cos\beta - r^2 ] ~, \label{ZRBSPK} \\ \tilde{Z}_I &=& {1\over 2} \, \tilde{\cal P}_{tu}^2 [ - \sin2\beta - 2r \cos\tilde{\Delta}\sin\beta ] ~, \label{ZIBSPK} \\ X &=& {\cal P}_{tu}^2 [ 1 + 2 r \cos{\Delta} \cos\beta + r^2 ] ~, \label{XBDKK} \\ Y &=& {\cal P}_{tu}^2 [ - 2 r \sin{\Delta} \sin\beta ] ~, \label{YBDKK} \\ Z &=& {\cal P}_{tu}^2 [ - \sin( 2\tilde{\beta} - 2\beta ) - 2 r \cos\Delta \sin( 2\tilde{\beta} - \beta ) - r^2 \sin2\tilde{\beta } ] ~. \label{ZBDKK} \end{eqnarray} Assuming that ${\tilde\beta}$ is measured in $B_d^0(t) \to J/\psi K_{\sss S}$, we now have six independent equations in six unknowns. We can solve for $\beta$ as follows. First, ${\cal P}_{tu} $ and $\tilde{\cal P}_{tu}$ are eliminated by dividing the equations as follows: \begin{eqnarray} \tilde{M} &\equiv & - \frac{\tilde{Z}_R}{\tilde{X}} = \frac{ \cos2\beta + 2r \cos{\tilde\Delta}\cos\beta + r^2 } { 1 + 2 r \cos{\tilde\Delta}\cos\beta + r^2 } = - 1 + \frac{2 \sin^2\beta }{ 1 + 2 r \cos{\tilde\Delta} \cos\beta + r^2 } \label{TMSM} \\ \tilde{N} &\equiv & \frac{\tilde{Z}_I}{\tilde{X}} = \frac{ \sin2\beta + 2r \cos{\tilde\Delta}\sin\beta } { 1 + 2 r \cos{\tilde\Delta}\cos\beta + r^2 } = - 2 \sin\beta \frac{\cos\beta + r \cos\tilde{\delta } } { 1 + 2 r \cos{\tilde\Delta}\cos\beta + r^2 } \label{TNSM} \\ \tilde{O} &\equiv & \frac{\tilde{Y}}{\tilde{X}} = \frac{ - 2 r \sin{\tilde\Delta} \sin\beta } { 1 + 2 r \cos{\tilde\Delta}\cos\beta + r^2 } \\ M &\equiv & \frac{{Z}}{{X}} + \sin2\tilde{\beta} =\sin2\tilde{\beta} ( \frac{ 2 \sin^2\beta } { 1 + 2 r \cos{\Delta}\cos\beta + r^2 } ) + \cos2\tilde{\beta} \frac{\sin2\beta + 2 r \cos\Delta \sin\beta } { 1 + 2 r \cos{\Delta}\cos\beta + r^2 } \label{M}\\ O &\equiv & \frac{{Y}}{{X}} = \frac{ - 2 r \sin{\Delta} \sin\beta } { 1 + 2 r \cos{\Delta}\cos\beta + r^2 } \label{O} \end{eqnarray} We then define \begin{eqnarray} R &\equiv & - \frac{\tilde{M}+1}{\tilde{N}} = \frac{\sin\beta}{\cos\beta + r \cos{\tilde\Delta}} \label{R} \end{eqnarray} Eliminating $\tilde{\delta}$ from Eqs.~(\ref{R}) and (\ref{TMSM}), we then find $r^2 $ as a function of $\beta $: \begin{eqnarray} r^2 = -1 + 2 \cos^2\beta + \frac{2}{\tilde{M}+1}\sin^2\beta - \frac{2}{R}\sin\beta \cos\beta \label{r} \end{eqnarray} And from Eq.~(\ref{M}) we have \begin{eqnarray} \cos\Delta = \frac{ -2 \sin2\tilde{\beta} \sin^2\beta -2 \cos2\tilde{\beta} \sin\beta\cos\beta + M + r^2 M } { 2r (- M \cos\beta + \cos2\tilde{\beta}\sin\beta )} \label{Cdel} \end{eqnarray} Finally, by inserting the expressions for $\Delta$ and $r^2 $ into Eq.~(\ref{O}) we obtain an equation for $\beta$ in terms of observables alone. Note that we have not used the expression for $\tilde{Y}$ [Eq.~(\ref{YBSPK})] in the above derivation. As explained earlier, $\tilde{Y}$ is not independent of $\tilde{X}$, $\tilde{Z}_R$ and $\tilde{Z}_I$. However, it can be used as a check to eliminate some of the solutions, thereby reducing the discrete ambiguity. We illustrate this solution numerically in Table~\ref{betatable}. By choosing input values for the theoretical parameters, we can generate the ``experimental data'' of Eqs.~(\ref{XBSPK})-(\ref{ZBDKK}). The above method can then be used to solve for the theoretical unknowns, and we can check that, despite the presence of multiple solutions, we can still find that $\beta \ne {\tilde\beta}$. For the amplitudes, we take $(\tilde{\cal P}_{tu})_{in}=1.0$ and $({\cal P}_{tu})_{in}=1.2$. (We take these quantities to be unequal in order to account for two things: (i) the different spectator quarks and (ii) the different final state -- two pseudoscalars in one case, and one vector and one pseudoscalar in the other case.) The assumed input values of $r$ and the weak and strong phases are shown in the Table. The angles are taken to lie in the region $0 < \beta, \Delta, {\tilde\Delta} < \pi $. \begin{table} \begin{center} \begin{tabular}{|r|r r r r|r|r|r|r|r|} \hline $\tilde{\beta}_{in}$ & $\beta_{in}$ & $r_{in}$ & $\tilde{\Delta }_{in} $ & $\Delta_{in} $ & $\beta $ & $ r $ & ${\tilde\Delta} $ & $\Delta $ & $ \tilde{P}_{tu}$ \\ \hline 25 & 10 & 0.3 & 80 & 120 & 19.7 & 1.22 & 28.1 & 180.0 & 0.26 \\ & & & & & 10.0 & 0.30 & 80.0 & 120.0 & 1.00 \\ & & & & & 160.3& 1.22 & 151.9 & 0.0 & 0.26 \\ & & & & & 170.0 & 0.30 & 100.0 & 60.0 & 1.00 \\ \hline 25 & 40 & 0.3 & 80 & 120 & 40.0 & 0.30 & 80.0 & 120.0 & 1.00 \\ & & & & & 57.2 & 0.65 & 35.8 & 165.9 & 0.59 \\ & & & & & 140.0 & 0.30 & 100.0 & 60.0 & 1.00 \\ & & & & & 122.8 & 0.65 & 144.2 & 14.1 & 0.59 \\ \hline 40 & 25 & 0.3 & 40 & 10 & 64.4 & 2.04 & 11.6 & 178.5 & 0.47 \\ & & & & & 25.0 & 0.30 & 40.0 & 10.0 & 1.00 \\ & & & & & 115.6 & 2,04 & 168.4 & 1.53 & 0.47 \\ & & & & & 155.0 & 0.30 & 168.3 & 170.0 & 1.0 \\ \hline \end{tabular} \end{center} \caption{Output values of $\beta$, $r$, ${\tilde\Delta}$, $\Delta$ and $\tilde{P}_{tu}$ for given input values of $r$ and the weak and strong phases. We take $(\tilde{\cal P}_{tu})_{in}=1.0$ and $({\cal P}_{tu})_{in}=1.2$. All phase angles are given in degrees.} \label{betatable} \end{table} {}From the Table, we see that $\beta$ can be extracted with a fourfold ambiguity. However, none of these values is equal to the value of ${\tilde\beta}$. Thus, if these measurements were carried out, and these results found, we would have unequivocal evidence of new physics in the $b\to d$ FCNC. We would not know if it affected $B_d^0$-${\overline{B_d^0}}$ mixing, the $b\to d$ penguin, or both, but we would know with certainty that new physics was present. In describing this method, we have assumed that there is no new physics in $B_s^0$-${\overline{B_s^0}}$ mixing. However, even if we include this, the above method does not change significantly. If a new-physics phase $\theta_s$ is present in $B_s^0$-${\overline{B_s^0}}$ mixing, then it is not the quantities ${\tilde Z}_R$ and ${\tilde Z}_I$ [Eqs.~(\ref{ZPR}) and (\ref{ZPI})] which are measured, but rather $\tilde{Z}_R^{ex}$ and $\tilde{Z}_I^{ex}$: \begin{eqnarray} \tilde{Z}_R^{ex} &\equiv & Re\left( e^{-i\theta_s }A_s^{\phi*} \bar{A}^{\phi}_{s } \right) = \cos\theta_s \tilde{Z}_I - \sin\theta_s \tilde{Z}_R ~, \\ \tilde{Z}_I^{ex} &\equiv & Im\left( e^{-i\theta_s } A_s^{\phi*} \bar{A}^{\phi}_{s } \right) = \cos\theta_s \tilde{Z}_R + \sin\theta_s \tilde{Z}_I ~. \end{eqnarray} However, ${\tilde Z}_R$ and ${\tilde Z}_I$ can be obtained straightforwardly: \begin{eqnarray} \tilde{Z}_R &=& \cos\theta_s \, \tilde{Z}_I^{ex} + \sin\theta_s \, \tilde{Z}_R^{ex} \\ \tilde{Z}_I &=& - \sin\theta_s \, \tilde{Z}_I^{ex} + \sin\theta_s \, \tilde{Z}_R^{ex} \end{eqnarray} Thus, assuming that $\theta_s$ is known independently (e.g.\ via any of the methods we have described earlier), we can use these expressions for ${\tilde Z}_R$ and ${\tilde Z}_I$ and simply apply the above method. If $\theta_s$ is only known up to a discrete ambiguity, then this simply increases the number of possible solutions for $\beta$. However, in general we will still be able to determine that $\beta \ne {\tilde\beta}$. Finally, we note that even if there is no new physics (i.e.\ $\beta = {\tilde\beta}$), this method still yields important information. If one probes $\beta$ in the conventional way via CP violation in $B_d^0(t) \to J/\PsiK_{\sss S}$, one extracts the function $\sin 2\beta$. This gives the angle $\beta$ up to a fourfold discrete ambiguity: if $\beta_0$ is the true solution, $\beta_0 + \pi$, ${\pi\over 2} - \beta_0$ and ${3\pi\over 2} - \beta_0$ are also solutions for $\beta$. Our technique can be used to eliminate two of these solutions. In particular, due to the presence of the $\sin\beta$ and $\cos\beta$ factors in Eqs.~(\ref{XBSPK})-(\ref{ZBDKK}), ${\pi\over 2} - \beta_0$ and ${3\pi\over 2} - \beta_0$ will not in general be among the solutions to these equations. However, $\beta_0 + \pi$ will still be allowed if we simultaneously take $\Delta \to \Delta + \pi$ and $\tilde\Delta \to \tilde\Delta + \pi$. Thus, in the absence of new physics, the above method can be used to reduce the discrete ambiguity in $\beta$ from a fourfold one to a twofold one. \section{$B_s^0(t)\to J/\Psi K_{\sss S}$ and $B_d^0(t)\to J/\Psi \pi^0$} It is not difficult to think of variations on the above method. As a second example, consider the decays $B_s^0\to J/\Psi K_{\sss S}$ and $B_d^0\to J/\Psi \pi^0$. Both of these decays get contributions from colour-suppressed tree diagrams, Zweig-suppressed gluonic penguins and electroweak penguins. If the penguins are not too small compared to the tree diagram, it may be possible to extract $\beta$ and compare it with ${\tilde\beta}$. The amplitudes for these decays can be written as \begin{eqnarray} A_s^{\psi} &\equiv & A\left(B_s\rightarrow J/\psi K_s \right) = \tilde{C} + \tilde{P}_1 + \tilde{P}_{\scriptscriptstyle EW} ~, \\ A_d^{\psi \pi } &\equiv & A\left(B_d\rightarrow J/\psi {\pi }^0 \right) = C + P_1 + P_{\scriptscriptstyle EW} ~, \end{eqnarray} where $\tilde{C}$ and ${C}$ are the colour-suppressed tree amplitudes in $B_s^0$ and $B_d^0$ decays, respectively. The combination of CKM matrix elements involved in these amplitudes is $V_{cb}^* V_{cd}$, which is real in the Wolfenstein parametrization. As before, we use CKM unitarity to eliminate the $u$-quark piece of the penguin contributions, allowing us to write \begin{eqnarray} A_s^{\psi} &\equiv & \tilde{\cal C}e^{i {\tilde\delta}_C} + \tilde{\cal P}_{1} e^{i {\tilde\delta}_{{P}}} e^{-i\beta } ~, \\ A_d^{\psi \pi} &\equiv & {\cal C}e^{i \delta_C} + {\cal P}_{1} e^{i \delta_{{P}}} e^{-i\beta } ~, \end{eqnarray} where \begin{eqnarray} \tilde{\cal C} e^{i {\tilde\delta}_C} & = & \tilde{C} + \tilde{P}_{1,c} - \tilde{P}_{1,u} ~, \nonumber\\ \tilde{\cal P}_1 e^{i {\tilde\delta}_P} e^{-i\beta } & = & \tilde{P}_{1,t} - \tilde{P}_{1,u} + \tilde{P}_{\scriptscriptstyle EW} \nonumber\\ {\cal C} e^{i \delta_C} & = & C + P_{1,c} - P_{1,u} ~, \nonumber\\ {\cal P}_1 e^{i {\delta}_P} e^{-i\beta } & = & P_{1,t} - P_{1,u} + P_{\scriptscriptstyle EW} ~. \label{CPdefs} \end{eqnarray} In terms of these quantities, the time-dependent measurements yield the following: \begin{eqnarray} \tilde{X} &\equiv & \frac{1}{2} \left( \left| A_s^{\psi}\right|^2 + \left| \bar{A}^{\psi}_s \right|^2 \right) = \tilde{\cal C}^2 +\tilde{\cal P}_1^2 + 2 \tilde{\cal C} \tilde{\cal P}_1 \cos{\tilde\delta} \cos\beta ~, \label{XBSJK} \\ \tilde{Y} &\equiv & \frac{1}{2} \left( \left| A_s^{\psi}\right|^2 - \left| \bar{A}^{\psi}_s \right|^2 \right) = - 2 \tilde{\cal C} \tilde{\cal P}_1 \sin{\tilde\delta} \sin\beta ~, \label{YBSJK} \\ \tilde{Z}_R &\equiv & Re\left( A_s^{\psi*} \bar{A}^{\psi}_{s } \right) = - \tilde{\cal C}^2 - \tilde{\cal P}_1^2 \cos2\beta - 2 \tilde{\cal C} \tilde{\cal P}_1 \cos{\tilde\delta} \cos\beta ~, \label{ZRBSJK} \\ \tilde{Z}_I &\equiv & Im\left( A_s^{\psi*} \bar{A}^{\psi}_{s } \right) = - \tilde{\cal P}_1^2 \sin2\beta - 2 \tilde{\cal C} \tilde{\cal P}_1 \cos{\tilde\delta} \sin\beta ~, \label{ZIBSJK} \\ {X} &\equiv & \frac{1}{2} \left( \left| A_d^{\psi \pi}\right|^2 + \left| \bar{A}^{\psi \pi }_d \right|^2 \right) = {\cal C }^{2} +{\cal P}_1^{2} + 2 {\cal C} {\cal P}_1 \cos\delta \cos\beta ~, \label{XBDJP} \\ {Y} &\equiv & \frac{1}{2} \left( \left| A_d^{\psi \pi}\right|^2 - \left| \bar{A}^{\psi \pi}_d \right|^2 \right) = - 2 {\cal C} {\cal P}_1 \sin\delta \sin\beta \label{YBDJP} \\ {Z} &\equiv & Im\left( e^{-2i \tilde{\beta }} A_d^{\psi \pi*} \bar{A}^{\psi \pi}_{d } \right) \nonumber \\ &~& \qquad\qquad = - {\cal C}^{2} \sin 2\tilde{\beta } - {\cal P}_1^{2} \sin(2 \tilde{\beta } - 2 \beta ) - 2 {\cal C} {\cal P}_1 \cos\delta \sin(2 \tilde{ \beta } - \beta ) ~, \label{ZBDJP} \end{eqnarray} where ${\tilde\delta} \equiv {\tilde\delta}_C - {\tilde\delta}_P$ and $\delta \equiv \delta_C - \delta_P$. Once again, this gives us six independent equations in seven unknowns. However, we can reduce the number of parameters by assuming that $r = \tilde{r}$, where $r \equiv {\cal P}_1 / {\cal C}$ and ${\tilde r} \equiv \tilde{\cal P}_1 / \tilde{\cal C}$. Looking at Eq.~(\ref{CPdefs}), it is clear that this assumption is well-justified -- in fact, within the spectator model, the equality is exact. In this case the observables become \begin{eqnarray} \tilde{X} &=& \tilde{C}^2 [ 1 + 2 r \cos{\tilde\delta} \cos\beta + r^2 ] \label{TXBSJK} \\ \tilde{Y} &=& \tilde{\cal C}^2 [ - 2 r \sin{\tilde\delta} \sin\beta ] \label{TYBSJK} \\ \tilde{Z}_R &=& \tilde{\cal C}^2 [ - 1 - 2r \cos{\tilde\delta}\cos\beta - r^2 \cos2\beta ] \label{TZRBSJK} \\ \tilde{Z}_I &=& \tilde{\cal C}^2 [ - 2r \cos{\tilde\delta}\sin\beta - r^2 \sin2\beta ] \label{TZIBSJK} \\ X &=& {\cal C}^2 [ 1 + 2 r \cos{\delta} \cos\beta + r^2 ] \label{XDJP} \\ Y &=& {\cal C}^2 [ - 2 r \sin{\delta} \sin\beta ] \label{YDJP} \\ Z &=& {\cal C}^2 [ - \sin2\tilde{\beta } - 2 r \cos\delta \sin( 2\tilde{\beta} - \beta ) - r^2 \sin( 2\tilde{\beta} - 2\beta ) ] \label{ZDJP} \end{eqnarray} The form of these equations is similar to that found for $B_s^0\to\phiK_{\sss S}$ and $B_d^0 \to K^0 {\overline{K^0}}$ decays [Eqs.~(\ref{XBSPK})-(\ref{ZBDKK})]. As in that case, assuming that ${\tilde\beta}$ is measured in $B_d^0(t) \to J/\psi K_{\sss S}$, we now have six independent equations in six unknowns. And as before, it is possible to solve these equations for the six parameters. We can thus obtain $\beta$, and test whether $\beta = {\tilde\beta}$ or not. It must be admitted, however, that from a theoretical point of view this method is much less compelling than the one which uses the decays $B_s^0\to\phiK_{\sss S}$ and $B_d^0 \to K^0 {\overline{K^0}}$. We have noted that the assumption that $r = \tilde{r}$ is well justified since the only difference between the decays $B_s^0\to J/\Psi K_{\sss S}$ and $B_d^0\to J/\Psi \pi^0$ is the flavour of the spectator quark. However, this is also problematic: if the flavour of the spectator quark is completely irrelevant, then we also have ${\cal C} = \tilde{\cal C}$ and $\delta = \tilde\delta$. But in this case the $B_d^0\to J/\Psi \pi^0$ decay gives us no extra information: we have $X = \tilde{X}$, $Y = \tilde{Y}$, and $Z = \tilde{Z}_R \cos 2{\tilde\beta} - \tilde{Z}_I \sin 2{\tilde\beta}$. So we are back to the situation of having three equations in four unknowns, which obviously cannot be solved. Thus, for this method to work, not only is it necessary for the penguin contributions to $B_s^0\to J/\Psi K_{\sss S}$ and $B_d^0\to J/\Psi \pi^0$ to be sizeable, there must also be significant differences in the sizes of the contributing amplitudes due to the flavour of the spectator quark. While it is possible that these conditions are fulfilled, it is theoretically disfavoured. Thus, the method involving the decays $B_s^0\to\phiK_{\sss S}$ and $B_d^0 \to K^0 {\overline{K^0}}$ is probably more promising. \section{Conclusions} CP-violating asymmetries in $B$ decays will be measured in the near future. If Nature is kind, we will find evidence of physics beyond the SM. The most obvious signal of new physics will be if the unitarity triangle constructed from measurements of the CP angles disagrees with that constructed from independent measurements of the sides. The problem here is that there are significant theoretical uncertainties in the measurements of the sides. Thus, even if there is a discrepancy, it may not provide compelling evidence for new physics -- it may simply be that the errors on the theoretical parameters have been underestimated. For this reason, it is important to find ways of directly probing for new physics in the $B$ system. Although there are several methods which will allow us to directly test for the presence of new physics in the $b \to s$ FCNC (either in $B_s^0$-${\overline{B_s^0}}$ mixing or in $b \to s$ penguins), finding new physics in the $b\to d$ FCNC ($B_d^0$-${\overline{B_d^0}}$ mixing or $b\to d$ penguins) is considerably more difficult. To date, no methods have been suggested which directly probe new physics in the $b\to d$ FCNC. In this paper we have discussed a method to search for new physics in the $b\to d$ FCNC. By making time-dependent measurements of the decays $B_s^0(t)\to\phiK_{\sss S}$ and $B_d^0(t) \to K^0 {\overline{K^0}}$ it is possible to compare the weak phase in $B_d^0$-${\overline{B_d^0}}$ mixing with that of the $t$-quark contribution to the $b\to d$ penguin. Since these phases are equal in the SM, any discrepancy would be a clear signal of new physics. The method is not entirely free of hadronic uncertainties: it does require some theoretical input. Still, the necessary assumption --- that a ratio of amplitudes in the two decays is equal --- is reasonably well-justified theoretically. We estimate the error in this assumption to be $\roughly< 20\%$. This method can be applied to other decay modes. For example, we have also examined the decays $B_s^0(t)\to J/\Psi K_{\sss S}$ and $B_d^0(t)\to J/\Psi \pi^0$. Although in principle new physics in the $b\to d$ FCNC can be found using this set of decays, the necessary conditions are theoretically disfavoured. Thus the decays $B_s^0(t)\to\phiK_{\sss S}$ and $B_d^0(t) \to K^0 {\overline{K^0}}$ are more promising. \section*{\bf Acknowledgments} C.S.K. wishes to acknowledge the financial support of 1997-sughak program of Korean Research Foundation. The work of D.L. was financially supported by NSERC of Canada and FCAR du Qu\'ebec. The work of T.Y. was supported in part by Grant-in-Aid for Scientific Research from the Ministry of Education, Science and Culture of Japan and in part by JSPS Research Fellowships for Young Scientists.
\section*{\S 1. Preliminaries} In what follows $\fS_{k}$ stands for the symmetric group on $k$ elements. Let $\lambda$ be a partition of the number $k$ and $t$ a $\lambda$-tableau. Recall that $t$ is called {\it standard} if the numbers in its rows and columns grow from left to right and downwards. Denote by $C_{t}$ the column stabilizer of $t$, ler $R_{t}$ be its row stabilizer. We further set $$ e_{t}=\mathop{\sum}\limits_{\tau\in C_{t};\;\sigma\in R_{t}}\eps(\tau)\sigma\tau, \quad \tilde e_{t}=\mathop{\sum}\limits_{\tau\in C_{t};\;\sigma\in R_{t}}\eps(\tau)\tau\sigma.\eqno{(0.1)} $$ Let $\Nee$ be the set of positive integers, $\bar\Nee$, another , ``odd'', copy of $\Nee$ and let $\Mee=\Nee\coprod\bar\Nee$ be ordered so that each element of the ``even'' copy, $\Nee$, is smaller than any element form the ``odd'' copy, while inside of each copy the order is the natural one. We will call the elemnts from $\Nee$ ``even'' and those form $\bar\Nee$ ``odd'' ones; so we can encounter an ``even'' odd element, etc. Let $I$ be the sequence of elements from $\Mee$ of length $k$. Let us fill in the tableau $t$ with elements from $I$ replacing element $\alpha$ with $i_{\alpha}$. The sequence $I$ is called $t$-{\it semistandard} if the elements of $t$ do not decrease from left to right and downwards; the ``even'' elements strictly increase along columns; the ``odd'' elements strictly increase along rows. The group $\fS_{k}$ naturally acts on sequences $I$. Let $\fA$ be the free supercommutative superalgebra with unit generated by $\{x_{i}\}_{i\in I}$. For any $\sigma\in\fS_{k}$ define $c(I, \sigma)=\pm 1$ from the equation $$ c(I, \sigma)x_{I}=x_{\sigma^{-1}I}\text{ where }x_{I}=x_{i_{1}}\dots x_{i_{k}}.\eqno{(0.2)} $$ Clearly, $c(I, \sigma)$ is a cocycle, i.e., $$ c(I, \sigma\tau)=c(\sigma^{-1}I, \tau)c(I, \sigma). $$ With the help of this cocycle a representation of $\fS_{k}$ in $T^{k}(V)=V^{\otimes k}$ for any superspace $V$ is defined: $$ \sigma v_{I}= c(I, \sigma^{-1})v_{\sigma I}, \text{ where $v_{I}= v_{i_{1}}\otimes \dots \otimes v_{i_{k}}$ and $v_{i_{\alpha}}\in V$ for each $\alpha$}.\eqno{(0.3)} $$ Let $\{v_{1}, \dots , v_{n}; v_{\bar 1}, \dots , v_{\bar m}\}$ be a basis of $V$ in the standard format (the even elements come first followed by the odd ones). Then the elements $v_{I}$ for all possible sequences $I$ of length $k$ and with elements from $$ R_V=\{1, \dots , n; \bar 1, \dots , \bar m\}\eqno{(0.4)} $$ form a basis of $T^{k}(V)$. The following theorem describes the decomposition of $T^{k}(V)$ into irreducible $\fS_{k}\times \fgl(V)$-modules. \ssbegin{1.1}{Theorem} {\em (Cf. \cite{S1}.)} The commutant of the natural $\fgl(V)$-action on $T^{k}(V)$ is isomorphic to $\Cee[\fS_{k}]$ and $$ T^{k}(V)=\mathop{\oplus}\limits_{\lambda: \lambda_{n+1}\leq m}S^{\lambda}\otimes V^{\lambda}, $$ where $S^{\lambda}$ is an irreducible $\fS_{k}$-module and $V^{\lambda}$ is an irreducible $\fgl(V)$-module. \end{Theorem} The following refinement of Theorem 1.1 holds: \ssbegin{1.2}{Theorem} If $t$ runs over the standard tableaux of type $\lambda$ and $I$ runs over semistandard $t$-sequences, then the family $\{e_{t}(v_{I})\}$ (resp. $\{\tilde e_{t}(v_{I})\}$) is a basis in $S^{\lambda}\otimes V^{\lambda}$. Moreover, for a fixed $t$ the families $\{e_{t}(v_{I})\}$ and $\{\tilde e_{t}(v_{I})\}$ span $V^{\lambda}$. \end{Theorem} Proof follows from results of \cite{S1}. \qed Let $U$ and $W$ be two superspaces with bases $u_{i}$ and $w_{j}$, where $i\in R_U$, $j\in R_W$, respectively, and where $$ R_U=\{1, \dots , k; \bar 1, \dots , \bar l\}\text{ and } R_W=\{1, \dots , p; \bar 1, \dots , \bar q\}. $$ The symmetric algebra $S^{\bcdot}(U\otimes W)$ is generated by $z_{ij}=u_{i}\otimes w_{j}$ for $i\in R_U$ and $j\in R_W$. Let $I$ be a sequence of length $N$ with elements from $R_U$ and $J$ a sequence of the same length with elements form $R_W$. Let $p(i_{\alpha})$ and $p(j_{\beta})$be the parities of the corresponding elements of the sequence. Set $\alpha(I, J)=\mathop{\sum}\limits_{\alpha >\beta}p(i_{\alpha})p(j_{\beta})$ and define an element of $S^{\bcdot} (U\otimes W)$ by setting $$ Z(I, J)=(-1)^{\alpha(I, J)}\mathop{\prod}\limits_{\alpha=1}^{N}Z_{i_{\alpha}j_{\beta}}. \eqno{(1.1)} $$ For a given tableau $t$ of order $N$ we define polynomials $$ \renewcommand{\arraystretch}{1.4} \begin{array}{l} P_{t}(I, J)=\mathop{\sum}\limits_{\sigma\in R_{t}, \, \tau\in C_{t}}\eps(\tau)c(I, (\sigma\tau)^{-1})Z(\sigma\tau I, J),\\ \tilde P_{t}(I, J)=\mathop{\sum}\limits_{\sigma\in R_{t}, \, \tau\in C_{t}}\eps(\tau)c(I, (\tau\sigma)^{-1})Z(\tau\sigma I, J). \end{array} $$ The Lie superalgebras $\fgl(U)$ and $\fgl(W)$ naturally act on $S^{\bcdot}(U\otimes W)$ and their actions commute. \ssbegin{1.3}{Theorem} $S^{\bcdot}(U\otimes W)=\mathop{\oplus}\limits_{\lambda}U^{\lambda}\otimes W^{\lambda}$, where $U^{\lambda}$ and $W^{\lambda}$ are irreducible $\fgl(U)$- and $\fgl(W)$-modules, respectively, corresponding to the partition $\lambda$ and the sum runs over partitions such that $\lambda_{\alpha +1}\leq \beta$ for $\alpha=\min(k, p)$ and $\beta=\min(l, q)$. \end{Theorem} \begin{proof} By Theorem 1.1 $$ W^{\otimes N}=\oplus ~ W^{\lambda}\otimes S^{\lambda}\text{ and } U^{\otimes N}=\oplus ~U^{\mu}\otimes S^{\mu}. $$ Hence, $$ \renewcommand{\arraystretch}{1.4} \begin{array}{l} S^{N}(U\otimes W)=((U\otimes W)^{\otimes N})^{\fS_{N}}= (U^{\otimes N}\otimes W^{\otimes N})^{\fS_{N}}=\\ \mathop{\oplus}\limits_{\lambda, \mu} (U^{\lambda}\otimes W^{\mu}\otimes S^{\lambda}\otimes S^{\mu})^{\fS_{N}}= \mathop{\oplus}\limits_{\lambda, \mu} (U^{\lambda}\otimes W^{\mu})\otimes (S^{\lambda}\otimes S^{\mu})^{\fS_{N}}. \end{array} $$ Since $(S^{\lambda})^{*}\simeq S^{\lambda}$ and since $S^{\lambda}$ and $S^{\mu}$ are irreducible, we have $$ (S^{\lambda}\otimes S^{\mu})^{\fS_{N}}=\Hom_{\fS_{N}}(S^{\lambda}, S^{\mu})=0 \text{ if $\lambda\neq \mu$ and $\Cee$ otherwise}. $$ Theorem is proved. \end{proof} \ssbegin{1.4}{Theorem} Let $t$ be a standard tableau of type $\lambda$ and let $I$ and $J$ be $t$-semistandard sequences. Then the family $P_{t}(I, J)$, as well as the similar family $\tilde P_{t}(I, J)$, forms a basis in the module $U^{\lambda}\otimes W^{\lambda}$. \end{Theorem} \begin{proof} The natural homomorphism $$ \phi_{N}: U^{\otimes N}\otimes W^{\otimes N}\tto S^{N}(U\otimes W) $$ is, clearly, a homomorphism of $\fgl(U)\oplus \fgl(W)$-modules. It is not difficult to verify that $$ \phi_{N}(e_{t}(v_{I})\otimes\tilde e_{t}(w_{J}))=c\cdot P_{t}(I, J)\text{ for a constant } c. $$ Let $t$ be a fixed $\lambda$-tableau, $I$ and $J$ two $t$-semistandard sequences with elements from $R_U$ and $R_W$, respectively. Then by Theorem 1.2 the vectors $e_{t}(v_{I})\otimes\tilde e_{t}(w_{J})$ form a basis of a subspace $L\subset U^{\otimes N}\otimes W^{\otimes N}$ which is also a $\fgl(U)\oplus \fgl(W)$-submodule. By the same theorem, $L\simeq U^{\lambda}\otimes W^{\lambda}$ and it remains to show that $\phi_{N}(L)\neq 0$. For this it suffices to show that there exists an $l\in L$ such that $\phi(l)\neq 0$. Since $\phi_{N}(\sigma v_{i}\otimes \sigma w_{J})= \phi_{N}(v_{I}\otimes w_{J})$, it follows that $$ \renewcommand{\arraystretch}{1.4} \begin{array}{l} \phi_{N}(e_{t}(v_{I})\otimes \tilde e_{t}(w_{J}))= c\phi_{N}(e_{t}(v_{I})\otimes w_{J})=\\ c\phi_{N}(\sigma e_{t}(v_{I})\otimes \sigma w_{J})= c\phi_{N}(e_{\sigma t}(\sigma v_{I})\otimes \sigma w_{J})=\\ \pm c\phi_{N}(e_{\sigma t}(v_{\sigma I})\otimes w_{\sigma J}). \end{array} $$ Therefore, we may assume that the tableau $t$ is consequtively filled in along the rows with the numbers 1, 2, etc. Observe that the sequences $\sigma I$ and $\sigma J$ remain $\sigma t$-semistandard. Let $I=J$ be the sequence $$ \underbrace{1\dots 1}_{\lambda_{1}} \underbrace{2\dots 2}_{\lambda_{2}}\dots \underbrace{\alpha\dots \alpha}_{\lambda_{\alpha}}\bar 1\dots \bar\lambda_{\alpha+1}\bar 1\dots \bar\lambda_{\alpha+2}\dots \bar 1\dots \bar\lambda_{\gamma}, $$ where $(\lambda_{1}, \lambda_{2}, \dots, \lambda_{\alpha}, \dots , \lambda_{\gamma})$ is the partition corresponding to $t$, $\alpha=\min(\dim~U_{\ev}, \dim~W_{\ev})$, $\beta=\min(\dim~U_{\od}, \dim~W_{\od})$. It is not difficult to verify that $\phi_{N}(e_{t}(v_{I})\otimes w_{I})\neq 0$. Since $\phi_{N}$ is a homomorphism of $\fgl(U)\oplus \fgl(W)$-modules, its restriction onto $L$ is an isomorphism. This implies the statement of Theorem for the family $P_{t}(I, J)$. For the family $\tilde P_{t}(I, J)$ proof is similar. \end{proof} Let us elucidate how the results obtained can be applied to the invariant theory. Let $\fg\subset \fgl(V)$ be a Lie superalgebra. Under ``the invariant theory of $\fg$'' we understand the description of $\fg$-invariants in the superalgebra $\fA^{p, q}_{k, l}=S(U\otimes V\; \oplus \; V^{*}\otimes W)$. On $\fA^{p, q}_{k, l}$, the Lie superalgebras $\fgl(U)$ and $\fgl(W)$ naturally act. By Theorem 1.3 we have $$ \fA^{p, q}_{k, l}=\mathop{\oplus}\limits_{\lambda, \mu} \; U^{\lambda}\otimes V^{\lambda}\otimes V^*{}^{\mu}\otimes W^{\mu}. $$ Therefore, to describe $\fg$-invariant elements, it suffices to describe the $\fg$-invariants in $V^{\lambda}\otimes V^*{}^{\mu}=\Hom(V^{\mu}, V^{\lambda})$. But $(V^{\lambda}\otimes V^*{}^{\mu})^\fg=\Hom_{\fg}(V^{\mu}, V^{\lambda})$, i.e., the description of $\fg$-invariants is equivalent to the description of $\fg$-homomorphisms of $\fg$-modules $V^{\mu}$. Let us consider how the method works in the simplest example: $\fg=\fgl(V)$. Let $\{e_{i}: i\in R_V\}$ be a basis of $V$ in a standard format; $\{e_{i}^{*}\}$ the left dual basis. Set $$ \theta=\mathop{\sum}\limits_{i\in T}e_{i}\otimes e_{i}^{*}, \quad \hat\theta=\mathop{\sum}\limits_{i\in T}(-1)^{p(i)}e_{i}^{*}\otimes e_{i}. $$ It is not difficult to verify that $\theta$ and $\hat\theta$ are $\fg$-invariants. Set $$ T^{p, q}(V)=V^{\otimes p}\otimes V^*{}^{\otimes q}, \quad \hat T^{p, q}(V)=V^*{}^{\otimes p}\otimes V^{\otimes q}. $$ On $T^{p, q}(V)$ and $\hat T^{p, q}(V)$, the group $\fS_{p}\times \fS_{q}$ acts and its action commutes with that fo $\fgl(V)$. Hence, $\fS_{p}\times \fS_{q}$ also acts on the space of $\fgl(V)$-invariants in $T^{p, q}(V)$ and $\hat T^{p, q}(V)$. \section*{\S 2. The invariants of \protect $\fgl(V)$} Set $v_r{}^*=(x_{r1}, \dots , x_{rn}; x_{r\bar 1}, \dots , x_{r\bar m})$ and $v_s=(x_{1s}{}^*, \dots , x_{n s}{}^*; x_{\bar 1s}{}^*, \dots , x_{\bar m s}{}^*)^{t}$, where $x_{ri}=u_r\otimes e_i$ and $x_{is}=e_i{}^*\otimes w_s$, i.e., $v_r{}^*$ is a row vector and $v_s$ is a column vector, so their scalar product is equal to $(v_r{}^*, v_s)=\mathop{\sum}\limits_i x_{ri}x_{i s}{}^*$. \ssbegin{2.1}{Theorem} The algebra of $\fgl(V)$-invariant elements in $\fA^{p, q}_{k ,l}$ is generated by the elements $(v_r{}^*, v_s)$ for all $r\in R_U$, $s\in R_W$. \end{Theorem} \begin{proof} Let $A$ be a supercommutative superalgebra, $L$ a $\fg$-module; let $L_A=(L\otimes A)_{\ev}$ and $\fg_A=(\fg\otimes A)_{\ev}$. The elements of $S^{\bcdot}(L^*)$ may be considered as functions on $L_A$ with values in $A$. Let $l\in L_A=(L\otimes A)_{\ev}=(\Hom(L^*, A))_{\ev}$. Therefore, $l$ determines a homomorphism $\phi_l: S(L^*)\tto A$. Set $$ f(l)=\phi_l(f)\text{ for any }f\in S^{\bcdot}(L^*). $$ Observe that $\fg_A$ naturally acts on $L_A$ and on the algebra of functions on $L_A$. \end{proof} \ssbegin{2.1.1}{Statement} {\em (\cite{S2})} Let $A$ be the Grassmann superalgebra with more indeterminates than $\dim L_\od$. The element $x$ of $S(L^*)$ is a $\fg$-invariant if and only if $x$ is a $\fg_A$-invariant when considered as a function on $L_A$. \end{Statement} \ssec{2.1.2. Proof of Theorem 2.1} Let $V^p$ denote $V\oplus \dots\oplus V$ ($p$ summands). Set $L=V^p\oplus \Pi(V)^q\oplus (V^*)^k\oplus \Pi(V^*)^l$; then $S(L^*)=\fA^{p, q}_{k, l}$ and we can consider $L_A$ as the set of collections $$ \cL=(v_1, \dots , v_p, v_{\bar 1}, \dots , v_{\bar q}, v_1^*, \dots , v_k^*, v_{\bar 1}^*, \dots , v_{\bar l}^*), $$ where $v_s\in V\otimes A$ and $v_t^*\in \Hom_A(V\otimes A, A)$ and the parity of these vectors coincide with the parities of their indices. Let us write the vectors with right coordinates and the covectors with left ones: $$ v_s=\mathop{\sum}\limits_i e_ia_{is}^*, \quad v_t^*=\mathop{\sum}\limits_i a_{ti}e_i^*. $$ Consider now the elements of $\fA^{p, q}_{k, l}$ as functions on $\cL$, by setting $$ x_{is}^*(\cL)=a_{is}^*, \quad x_{ti}^*(\cL)=a_{ti}. $$ Therefore, thanks to 2.1.1 it suffices to describe the functions on $\cL$ contained in the subalgebra generated by the coordinate functions {\bf ??} and invariants with respect to $GL(V\otimes A)$. Now, since the scalar products turn into scalar products under the $\fgl(U)\oplus \fgl(W)$-action, it is sufficient to confine ourselves to the invariants in $\fA^{n, m}_{n, m}$. Denote by $M$ the set of collections $(v_1, \dots , v_n, v_{\bar 1}, \dots , v_{\bar m})$ that form bases of $V\otimes A$. In Zariski topology the set $M$ is dense in the space of all collections. If $f$ is an invariant and $\cL\in M$, then there exists $g\in GL(V\otimes A)$ such that $gv_i=e_i$ for each $i\in T$. Therefore, $f(\cL)=f(g\cL)=f(e_1, \dots , e_n, gv_{\bar 1}^*, \dots , gv_{\bar m}^*)$ and $f(\cL)$ is a polynomial in coordinates of the $gv_{\bar t}$. But $(gv_{\bar t}^*, e_i)=(v_{\bar t}^*, g^{-1}e_i)=(v_{\bar t}^*, v_i)$ which proves the theorem. \qed \begin{Corollary} The nonzero $\fgl(V)$-invariants in $T^{p, q}$ only exist if $p=q$. In this case the $\fS_p\times \fS_p$-module of invariants is generated by the images of the canonical elements $\theta^{\otimes p}$ in $T^{p, p}$ and $\hat\theta$ in $\hat T^{p, p}$. \end{Corollary} Consider now the algebra homomorphism $$ S^{\bcdot}(U\otimes W)\tto (\fA^{p, q}_{k, l})^{\fgl(V)}, \quad u_{r}\otimes w_s\mapsto (v_r^*, v_s).\eqno{(2.1)} $$ The kernel of this homomorphism is the ideal of relations between the scalar products. \ssbegin{2.2}{Theorem} The ideal of relations between scalar products $(v_r^*, v_s)$ is generated by the polynomials $P_t(I, J)$, where $t$ is a fixed standard rectangular $(n+1)\times (m+1)$ tableau, $I$ and $J$ are $t$-semistandard sequences with elements from $R_U$ and $R_W$, respectively. \end{Theorem} \begin{proof} By Theorem 1.3 $S^{\bcdot}(U\otimes W)=\mathop{\oplus}\limits_{\lambda}U^\lambda\otimes W^\lambda$ and $$ \fA^{p, q}_{k, l}=S^{\bcdot}(U\otimes V\oplus V^*\otimes W)=S^{\bcdot}(U\otimes V)\otimes S^{\bcdot}(V^*\otimes W)=(\mathop{\oplus}\limits_{\mu}U^\mu\otimes V^\mu)\otimes (\mathop{\oplus}\limits_{\nu}(V^*)^\nu\otimes W^\nu), $$ hence, $(\fA^{p, q}_{k, l})^{\fgl(V)}=\mathop{\oplus}\limits_{\mu: \mu_{n+1}\leq m}U^\mu\otimes W^\mu$. Since homomorphism (2.1) is a homomorphism of $\fgl(V)\oplus \fgl(W)$-modules, its kernel coinsides with $\mathop{\oplus}\limits_{\lambda: \lambda_{n+1}\geq m+1}U^\lambda\otimes W^\lambda$. Let $\nu$ be a $(n+1)\times (m+1)$ rectangle. The condition $\lambda_{n+1}\geq m+1$ means that $\lambda\supset \nu$ and by Theorem 1.3 it suffices to demonstrate that $P_t(I, J)$, where $t$ is a fixed standard rectangular tableau of size $\lambda$, belongs to the ideal generated by $U^\nu\otimes W^\nu$. Let $e_t$ be the corresponding minimal idempotent, $e_s$ the minimal idempotent for a standard tableau $s$ of size $\nu$. Decomposing $R_t$ with respect to the right cosets relative $R_s$ and decomposing $C_t$ with respect to the left cosets relative $C_s$ we obtain a representation of $e_t$ in the form $\sum\tau_ie_s\sigma_j$. This implies that $P_t(I, J)$ is the sum of polynomials of the form $f_iP_{t_{i}}(I_{i}, J_{j})\phi_{j}$, i.e., belongs to the ideal generated by the $P_t(I, J)$. \end{proof} \section*{\S 3. The invariants of \protect $\fsl(V)$} First, let us describe certain tensor invariants. Obviously, all $\fgl(V)$-invariants are also $\fsl(V)$-invariants. Therefore, we will only describe the $\fsl(V)$-invariants which are not $\fgl(V)$-invariants. Denote by $\theta_k=\theta^{\otimes k}$ the invariant in $T^{k, k}$ and by $\hat\theta_k=\hat\theta^{\otimes k}$ the invariant in $\hat T^{k, k}$, and for a given sequence $I$ with elements from $R_V$ set $$ v_I=e_{i_{1}}\otimes \dots \otimes e_{i_{k}}\text{ and } v_I^*=e_{i_{1}}^*\otimes \dots \otimes e_{i_{k}^*}. $$ Let us represent $\fsl(V)$ in the form $\fg=\fg_-\oplus\fg_0\oplus\fg_+$, where $\fg_0=\fg_\ev$ and $\fg_\pm$ are the $\fg_0$-modules generated by the positive and negative root vectors, respectively. Let $\{X_\alpha\}_{\alpha\in R^-}$ and $\{X_\beta\}_{\beta\in R^+}$, where $R^{\pm}$ are the sets of positive (negative) roots, be some bases of $\fg_-$ and $\fg_+$, respectively; set $X_-=\prod X_\alpha$ and $X_+=\prod X_\beta$. The elements $X_\pm$ are uniquely determined up to a constant factor because the subalgebras $\fg_\pm$ are commutative. \ssbegin{3.1}{Lemma} Let $M$ be a $\fg_0$-module and $\tilde M=\ind^{\fg}_{\fg_0}(M)$ be the induced $\fg$-module. Then each of the correspondences $m\mapsto X_+X_-m$ and $m\mapsto X_-X_+m$ is a bijection of $M^{\fg_{0}}$ onto $\tilde M^{\fg}$. \end{Lemma} \begin{proof} As follows from Lemma 4.2 below, $\dim M^{\fg_{0}}=\dim \tilde M^{\fg}$. Therefore, it suffices to show that the correspondence $m\mapsto n=X_+X_-m$ is injective map of $M^{\fg_{0}}$ to $\tilde M^{\fg}$. The injectivity is manifest, so we only have to check that the image is $\fg$-invariant. Clearly, $\fg_+n=\fg_0n=0$. Therefore, it suffices to verify that $X_{-\alpha}n=0$ for every simple root $\alpha$. This is subject to a direct check with the help of the multiplication table in $\fgl(V)$. \end{proof} \ssbegin{3.2}{Lemma} Let $V_1$ and $V_2$ be finite dimensional $\fg_0$-modules. Set $\fg_+V_1=\fg_-V_2=0$. Then $$ \ind^{\fg}_{\fg_0\oplus\fg_+}V_1 \otimes\ind^{\fg}_{\fg_0\oplus\fg_-}(V_2) \simeq \ind^{\fg}_{\fg_0}(V_1\otimes V_2).\eqno{(3.2.1)} $$ is an isomorphism of $\fg$-modules. \end{Lemma} \begin{proof} Since the dimensions of both modules are equal, it suffices to show that the natural homomorphism $$ \ind^{\fg}_{\fg_0}(V_1\otimes V_2)\tto \ind^{\fg}_{\fg_0\oplus\fg_+}V_1 \otimes\ind^{\fg}_{\fg_0\oplus\fg_-}(V_2).\eqno{(3.2.2)} $$ is surjective, i.e., the module generated by $V_1\otimes V_2$ coinsides with the whole module. The module in the right hand side has a natural filtration induced by filtrations of the modules $\ind^{\fg}_{\fg_0\oplus\fg_+}V_1$ and $\ind^{\fg}_{\fg_0\oplus\fg_-}(V_2)$. Let the $X_\alpha$ be a basisi of $\fg_+$ and $X_{-\alpha}$ a basisi of $\fg_-$. Consider the module $W$ generated by $V_1\otimes V_2$, i.e., by the elements of filtration zero and the element $$ w=X_{-\alpha_{1}}\dots X_{-\alpha_{k}}v_1\otimes X_{\beta_{1}}\dots X_{\beta_{l}}v_2. $$ We have $$ \renewcommand{\arraystretch}{1.4} \begin{array}{l} w=X_{-\alpha_{1}}(X_{-\alpha_{2}}\dots X_{-\alpha_{k}}v_1\otimes X_{\beta_{1}}\dots X_{\beta_{l}}v_2) \pm X_{-\alpha_{2}}\dots X_{-\alpha_{k}}v_1\otimes X_{-\alpha_{1}}X_{\beta_{1}}\dots X_{\beta_{l}}v_2=\\ X_{-\alpha_{1}}(X_{-\alpha_{2}}\dots X_{-\alpha_{k}}v_1\otimes X_{\beta_{1}}\dots X_{\beta_{l}}v_2)\pm \\ \pm X_{-\alpha_{2}}\dots X_{-\alpha_{k}}v_1\otimes (\mathop{\sum}\limits_i X_{\beta_{1}}\dots X_{\beta_{i-1}} [X_{-\alpha_{1}}, X_{\beta_{i}}]X_{\beta_{i+1}}\dots X_{\beta_{l}}v_2. \end{array} $$ Since each summand is of filtration $<k+l$, they belong to $W$ by inductive hypothesis; hence, so does $w\in W$. \end{proof} Let $t$ be a tableau consisting of $m$ columns and $n+k$ rows and filled in as follows: first, we fill in the tableau $t_1$ that occupies the first $n$ rows, next, the tableau $t_2$ that occupies the remaining rows, both tableaux are filled in consequtevely column-wise. Let $s$ be a tableau consisting of $n$ rows and $k+m$ columns and filled in as follows: first, we fill in the tableau $s_1$ that occupies the first $k$ columns, next, the tableau $t_2$ that occupies the remaining columns, both tableaux are filled in consequtevely column-wise. Let $I_k$ be the sequence obtained by $k$-fold repetition of the sequence $1, 2, \dots , n$; let $J_k$ be the sequence consisting of $k$ copies of $\bar 1$ in a row, next $k$ copies of $\bar 2$ in a row, etc., $k$ copies of $\bar m$ in a row. \ssbegin{3.3}{Theorem} In $\hat T^{(m+k)n, (n+k)m}(V)$, the element $$ e_s\times \tilde e_t(v^*_{I_{k}}\otimes\hat\theta_{nm}\otimes v_{J_{k}}) $$ is an $\fsl(V)$-invariant. \end{Theorem} \begin{proof} Let $N$ be any positive integer. Consider the map $\phi: T^{N, N}(V)\tto T^{N, N}(V_\ev)$ such that $\phi(V_\od)=\phi(V^*_\od)=0$. Clearly, $\phi$ is an $\fS_N\times \fS_N$-module homomorphism because it is induced by projections of $V$ and $V^*$ onto their even parts. Take $X_+$ and $X_-$ from Lemma 3.1 and consider the map $$ \psi: T^{N, N}(V_\ev)\tto T^{N, N}(V),\quad v_0\mapsto X_+X_-v_0. $$ Clearly, $\psi$ is an $\fS_N\times \fS_N$-module homomorphism. Let us consider the restrictions of the maps $\phi$ and $\psi$ onto $T^{N, N}(V)^{\fgl(V)}$ and $T^{N, N}(V_\ev)^{\fgl(V_\ev)}$, respectively. Clearly, $\phi$ sends the first of these spaces into the second one, whereas by Lemma 3.1 $\psi$ sends the second of these spaces into the first one. Theorem 1.1 implies that, as $\fS_N\times \fS_N$-modules, the spaces $T^{N, N}(V)^{\fgl(V)}$ and $T^{N, N}(V_\ev)^{\fgl(V_\ev)}$ have simple spectra. Let $S^\lambda\otimes S^\lambda\subset T^{N, N}(V)^{\fgl(V)}$ while $S_0^\lambda\otimes S_0^\lambda\subset T^{N, N}(V_\ev)^{\fgl(V_\ev)}$ correspond to a typical diagram $\lambda$ and both are nonzero, i.e., $\lambda_n\geq m$ and $\lambda_{n+1} =0$. Then the simplicity of the spectrum and Lemma 3.1 imply that $\phi$ and $\psi$ are, up to a constant factor, mutually inverse isomorphisms of the modules $S^\lambda\otimes S^\lambda$ and $S_0^\lambda\otimes S_0^\lambda$. Let $$ C_t=\mathop{\cup}\limits_\tau\tau(C_{t_{1}}\times C_{t_{2}}),\quad R_s=\mathop{\cup}\limits_\sigma\sigma(R_{s_{1}}\times R_{s_{2}}) $$ be the decomposition of the column stabilisor $C_t$ of the tableau $t$ with respect to the left cosets relative the product of the column stabilisors of $t_1$ and $t_2$ and same of the row stabilisor $R_s$. Then $$ \tilde e_t=\mathop{\sum}\limits_\tau\eps(\tau)\tau\tilde e_{t_{1}}\tilde e_{t_{2}},\quad e_s=\mathop{\sum}\limits_\sigma \sigma e_{s_{1}}e_{s_{2}}. $$ It is easy to verify that $$ \fg_+(V_\ev)=\fg_-(V_\ev ^*)=\fg_+(V_\od ^*)=\fg_-(V_\od)=0.\eqno{(3.3)} $$ The vector $X_+e_s(v^*_{I_{k}}\otimes v^*_{I_{m}})$ belongs to a typical module, is a highest one with respect to $\fg_+\oplus(\fg_0)_+$ and nonzero, where $(\fg_0)_+$ is the set of strictly upper-triangular matrices with respect to the fixed basis of $V_\ev$. But (0.2) implies that $e_s(v^*_{I_{k}}\otimes v^*_{J_{n}})$ is also highest with respect to $\fg_+\oplus(\fg_0)_+$ and lies in the same module. This shows that $$ X_+e_s(v^*_{I_{k}}\otimes v^*_{I_{m}})=c\cdot e_s(v^*_{I_{k}}\otimes v^*_{J_{n}}),\text{ where }c\neq 0. $$ Further, from Lemmas 3.1 and 3.2 it follows that the vector $$ X_-X_+\left[e_s(v^*_{I_{k}}\otimes v^*_{I_{m}})\otimes\tilde e_t(v_{I_{m}}\otimes v_{J_{k}})\right] $$ is $\fg$-invariant because $e_s(v^*_{I_{k}}\otimes v^*_{I_{m}})\otimes\tilde e_t(v_{I_{m}}\otimes v_{J_{k}})$ is $\fg_{\ev}$-invariant. We make use of the fact that $X_+\tilde e_t(v^*_{I_{m}}\otimes v^*_{J_{k}})=0$ to decuce that $$ \renewcommand{\arraystretch}{1.4} \begin{array}{l} w=X_-X_+\left[e_s(v^*_{I_{k}}\otimes v^*_{I_{m}})\otimes\tilde e_t(v_{I_{m}}\otimes v_{J_{k}})\right]= X_-\left[[X_+e_s(v^*_{I_{k}}\otimes v^*_{I_{m}})]\otimes\tilde e_t(v_{I_{m}}\otimes v_{J_{k}})\right]=\\ const\cdot X_-[e_s(v^*_{I_{k}}\otimes v^*_{J_{n}})\otimes\tilde e_t(v_{I_{m}}\otimes v_{J_{k}})]=\\ \mathop{\sum}\limits_{\sigma, \tau}\eps(\tau)\sigma\times \tau\left (X_-[e_{s_{1}}e_{s_{2}}(v^*_{I_{k}}\otimes v^*_{I_{m}})\otimes\tilde e_{t_{1}}\tilde e_{t_{2}}(v_{I_{m}}\otimes v_{J_{k}})]\right)=\\ \mathop{\sum}\limits_{\sigma, \tau}\eps(\tau)\sigma\times \tau\left (e_{s_{1}}(v^*_{I_{k}})\otimes X_-[e_{s_{2}}v^*_{I_{m}})\otimes\tilde e_{t_{1}}(v_{I_{m}})]\tilde e_{t_{2}}(v_{J_{k}})\right)=\\ \mathop{\sum}\limits_{\sigma, \tau}\eps(\tau)\sigma\times \tau\left (e_{s_{1}}(v^*_{I_{k}})\otimes X_-X_+[e_{s_{2}}(v^*_{I_{m}})\otimes\tilde e_{t_{1}}(v_{I_{m}})]\tilde e_{t_{2}}(v_{J_{k}})\right). \end{array} $$ Further on, $$ \phi(e_{s_{2}}\times\tilde e_{t_{1}}(\hat\theta_{nm}))= e_{s_{2}}\times\tilde e_{t_{1}}(\phi(\hat\theta_{nm}))= e_{s_{2}}\times\tilde e_{t_{1}}(\sum v^*_L\otimes v_L), $$ where $L$ runs over all the sequences of length $nm$ composed from the integers 1 to $n$. But, as is not difficult to see, $$ e_{s_{2}}\times\tilde e_{t_{1}}(\sum v^*_L\otimes v_L)= const\cdot e_{s_{2}}(v^*_{I_{m}})\otimes\tilde e_{t_{1}}(v_{I_{m}}), $$ hence, $$ X_-X_+e_{s_{2}}(v^*_{I_{m}})\otimes\tilde e_{t_{1}}(v_{I_{m}})=const\cdot e_{s_{2}}\times e_{s_{1}} (\hat\theta_{nm}). $$ Therefore, $$ w=const\cdot \mathop{\sum}\limits_{\sigma, \tau}\eps(\tau)\sigma\times \tau\left (e_{s_{1}}(v^*_{I_{k}})\otimes e_{s_{2}}\times\tilde e_{t_{1}}(\hat\theta_{nm})\otimes e_{t_{2}}(v_{J})\right)= e_{s}\times\tilde e_{t}(v^*_{I_{k}}\otimes\hat\theta_{nm} \otimes v_{J}) $$ which proves the theorem. \end{proof} Proof of the following theorem is similar. \ssbegin{3.4}{Theorem} The element $e_{s}\times\tilde e_{t}(v_{I_{k}}\otimes\theta_{nm}\otimes v_{J_{k}}^*)$ in $T^{nm+kn, nm+km}(V)$ is $\fsl(V)$-invariant. \end{Theorem} \ssbegin{3.5}{Corollary} Let $L$ eb the sequence with elements from $\Mee$. Set $$ p(L)=\sum p(l_{i}),\quad \alpha(L, L)=\mathop{\sum}\limits_{i<j}p(l_{i})p(l_{j}). $$ Under notations of Theorems $3.3$, $3.4$ the invariant elements can be expressed in the form $$ e_{s}\times\tilde e_{t}(v^*_{I_{k}}\otimes\theta_{nm}\otimes v_{J_{k}})= \mathop{\sum}\limits_L(-1)^{p(L)+\alpha(L, L)}e_{s}(v^*_{I_{k}}\otimes v^*_L)\otimes\tilde e_{t}(v_L\otimes v_{J_{k}})\eqno{(3.5.1)} $$ and $$ e_{s}\times\tilde e_{t}(v_{I_{k}}\otimes\theta_{nm}^*\otimes v^*_{J_{k}})= \mathop{\sum}\limits_L(-1)^{\alpha(L, L)}e_{s}(v_{I_{k}}\otimes v_L)\otimes\tilde e_{t}(v^*_L\otimes v^*_{J_{k}}),\eqno{(3.5.2)} $$ where the sums run over all the sequences $L$ of length nm with elements from $R_V$. \end{Corollary} \begin{proof} It is easy to verify that $\hat \theta_{nm}=\mathop{\sum}\limits_L (-1)^{\alpha(L, L)+p(L)}v^*_L\otimes v_L$, which immediately implies (3.5.1). Formula (3.5.2) is similarly proved. \end{proof} \ssec{3.6} Recall the definition of $R_V$, $R_U$ and $R_W$ (0.4) and (0.5). For any sequences $I$ and $J$ denote by $I*J$ the sequence obtained by ascribing $J$ at the end of $I$. Let now $I$ be the sequence of length $(k+m)n$ with elements from $R_U$, let $J$ be the sequence of length $(k+n)m$ with elements from $R_W$, let $\hat I$ be the sequence of length $(k+n)m$ with elements from $R_U$ and $\hat J$ be the sequence of length $(k+m)n$ with elements from $R_W$. For any sequence $L$ of length $nm$ with elements from $R_V$ we define: $$ \renewcommand{\arraystretch}{1.4} \begin{array}{l} \tilde P_s(I, I_k*L)\in S^{\bcdot}(U\otimes V),\quad \tilde P_t(L*J_k, J)\in S^{\bcdot}(V^*\otimes W);\\ P_t(\hat I, L*J_k)\in S^{\bcdot}(U\otimes V), \quad P_s(I_k*L, \hat J)\in S^{\bcdot}(V^*\otimes W). \end{array} $$ \begin{Theorem} The algebra of $\fsl(V)$-invariant elements in $\fA^{p, q}_{k, l}$ is generated by the elements {\em i)} $(v^*_r, v_s)$, where $r\in R_U$ and $s\in R_W$; {\em ii)} $F_k(I, J)=\mathop{\sum}\limits_L(-1)^{\alpha(L, L)}\tilde P_s(I, I_k*L)\tilde P_t(L*J_k, J)$, where $I$ is an $s$-semistandard sequence and $J$ is a $t$-semistandard one; {\em iii)} $F_{-k}(\hat I, \hat J)= \mathop{\sum}\limits_L(-1)^{\alpha(L, L)+p(L)(p(\hat I)+p(\hat J))} P_s(I_k*L, \hat J) P_t(\hat I, L*J_k)$, \noindent where $\hat I$ is an $s$-semistandard sequence, $\hat J$ is a $t$-semistandard one and $L$ runs over all the sequences of length $nm$ with elements from $R_V$. \end{Theorem} \begin{proof} For Young tableaux $\lambda$ and $\mu$ we have $$ (V^\lambda\otimes V^*{}^\mu)^{\fsl(V)}=\Hom_{\fsl(V)}(V^\mu, V^\lambda). $$ The dimension of this space is equal to either 0 or 1. It is equal to 1 only if $\lambda=\mu$ or both of them contain a $n\times m$ rectangle and $\lambda_i=\mu_i+k$ for $i= 1, \dots , n$ and $\lambda'_j=\mu'_j+k$ for $j=1, \dots , m$ and any $k\in \Zee$. To prove the theorem, it suffices to show that for the above $\lambda$ and $\mu$ the module $V^\lambda\otimes V^*{}^\mu$ containes an invariant which can be expressed via the invariants listed in the theorem. By \cite{S2} such an invariant exists. Under the canonical homomorphism of the tensor algebra onto the symmetric one, the invariants of the form i)--iii) turn into a system of generators. Theorem is proved. \end{proof} \ssec{3.7} To the invariant element in $T^{n(m+k), m(n+k)}(V)$ there corresponds an invariant operator $T^{m(n+k)}(V)\tto T^{n(m+k)}(V)$. To describe it, observe that $C_t$ can be represented as $C_t=\mathop{\coprod}\limits_{\pi\in Z}(C_{t_{1}}\times C_{t_{2}})\pi$, the decomposition into right cosets relative the product of the column stabilizers of tableaux $t_{1}$ and $t_{2}$; let $Z$ be a collection of their representatives. Define $$ D_{J_{k}}: T^{m(n+k)}(V)\tto T^{nm}(V),\quad D_{J_{k}}(v_1\otimes v_2)=(-1)^{p(J_{k})p(v_{1})} v_1\cdot v^*_{J_{k}}(v_2). $$ \begin{Lemma} Let $\cL$ be an invariant operator corresponding to $e_{s}\times\tilde e_{t}(v_{I_{k}}\otimes\theta_{nm}\otimes v^*_{J_{k}})$. Then $$ \cL(e_t(v_L))=const\cdot \cL(v_L)=e_{s}(v_{I_{k}}\otimes D^*{}_{J_{k}}e_{t_{2}}\mathop{\sum}\limits_{\pi\in Z}\eps(\pi)\pi v_L).\eqno{(3.7)} $$ \end{Lemma} \begin{proof} To $\theta_{nm}$ there corresponds the identity operator $\id: V^{\otimes nm}\tto V^{\otimes nm}$; hence, to $\theta_{nm}\otimes v^*_{J_{k}}$ there corresponds the operator $D_{J_{k}}: V^{\otimes m(n+k)}\tto V^{\otimes nm}$ and to $v_{I_{k}}\otimes\theta_{nm}\otimes v^*_{J_{k}}$ there corresponds the operator $v_{I_{k}}\otimes D_{J_{k}}$; finally, to $e_{s}\times\tilde e_{t}(v_{I_{k}}\otimes\theta_{nm}\otimes v^*_{J_{k}})$ there corresponds the operator $e_{s}(v_{I_{k}}\otimes D_{J_{k}})e_{t}$. Hence, $$ \renewcommand{\arraystretch}{1.4} \begin{array}{l} \cL(e_{t}(v_L))=e_{s}v_{I_{k}}\otimes D_{J_{k}})e_{t}^ 2(v_L= c_{1}e_{s}v_{I_{k}}\otimes D_{J_{k}})e_{t}(v_L)= c_{1}\cdot \cL(v_L)=\\ c_{1}e_{s}(v_{I_{k}}\otimes D_{J_{k}}\mathop{\sum}\limits_\pi e_{t_{1}}e_{t_{2}}\eps(\pi)\pi v_L)= c_{1}e_{s}(v_{I_{k}}\otimes e_{t_{1}}D_{J_{k}}e_{t_{2}}\mathop{\sum}\limits_\pi \eps(\pi)\pi v_L)=\\ c_{1}e_{s}e_{t_{1}}(v_{I_{k}}\otimes D_{J_{k}}e_{t_{2}}\mathop{\sum}\limits_\pi \eps(\pi)\pi v_L)= c_{1}c_{2}e_{s}(v_{I_{k}}\otimes D_{J_{k}}e_{t_{2}}\mathop{\sum}\limits_\pi \eps(\pi)\pi v_L). \end{array} $$ The last equality follows from $e_{s}e_{t_{1}}=c_{2}e_{s}$. \end{proof} \ssec{3.8} Let us consider the case $k=1$ in more detail. Let $L$ be a sequence of length $nm+m$ with elements from $R_V$, considered as a $t$-tableau. In each column $L$, mark an ``odd" element so that all the elements marked, say, $l=(l_1, \dots , l_m)$, are distinct. The pair $(L, l)$ will be called a {\it marked} tableau. Introduce the following notations: $c_i$ for the parity of the $i$-th column, $d_i$ for the parity of the last elelment in the $i$-th column, $b_i$ for the parity of the column under the $i$-th marked element, $|b_i|$ for the number of elements in the $i$-th column under the $i$-th marked element, $\eps(l)$ for the sign of the permutation $l=(l_1, \dots , l_m)$ and $\eps(L, l)=(-1)^{q(L)}\eps(l)$; set further $$ \eps(L)=c_2+c_4+\dots +d_2+d_4+\dots; \; \; q(L)=b_1+|b_1|+b_2+|b_2|+\dots . $$ \begin{Theorem} The invariant operator is of the form $$ \cL(e_{t}(v_L))=const\cdot \cL(v_L)=const\cdot\eps(L)\mathop{\sum}\limits_{(L, l)} \eps(L, l)e_{s}(v_{I_{1}}\otimes v_{L\setminus l}),\eqno{(3.8)} $$ where the constant factor does not depend on $L$.\end{Theorem} \begin{proof} Since for the representatives of the cosets of $\fS_{n+1}/\fS_{n}$ we can take a collection of cycles, we may assume in formula (3.7) that $$ \pi=\pi_1\dots\pi_m,\;\; \pi_i\pi_j=\pi_j\pi_i\text{ for any }i, j. $$ Hence, $\pi^ 2=1$. Furhter on, $D_{J_{1}} e_{t_{2}}\sum\eps(\pi)\pi v_{L}\neq 0$ if and only if the last row of $\pi L$ for some $\pi$ is, up to a permutation, a permutation of $\{\bar 1, \dots , \bar m\}$. The set of marked tableau $(L, l)$ is in one-to-one correspondence with the set of pairs $(L, \pi)$ such that the last row of $\pi L$ is, up to a permutation, $\{\bar 1, \dots , \bar m\}$. Indeed, from the pair $(L, l)$ determine $\pi=\pi_1\dots\pi_m$, where $\pi_i$ is the cycle that shifts the elements under the $i$-th marked one one cell up along the column and places the marked one at the bottom. If the marked element lies in the last row, we set $\pi_i=1$. And, the other way round, given $\pi$, we mark $\pi(k_1)$, \dots , $\pi(k_m)$, where $(k_1, \dots , k_m)$ is the last row of $L$. Hence, (3.7) implies that $$ \cL(v_L)=\mathop{\sum}\limits_{(L, l)} \delta(L, l)e_{s}(v_{I_{1}}\otimes v_{L\setminus l}), $$ where $\delta(L, l)$ is a sign depending on $(L, l)$. Direct calculations of this sign lead us to (3.8). \end{proof} \section*{\S 4. The absolute invariants of \protect $\fosp(V)$} Let $A=U(\fosp(V))[\eps]$ be the central extension with the only extra relation $\eps ^2=1$. On $A$, introduce the coalgebra structure making use of that on $U(\fosp(V))$ and setting $\eps\mapsto \eps\otimes \eps$. Assuming that $\eps$ acts on $V$ as the scalar operator of multiplication by $-1$, we may consider $V$ as an $A$-module. Using the coalgebra structure on $A$, one can determine a natural $A$-action in $T^{p, q}(V)$ and $\fA^{p, q}_{k, l}$. So, we can speak about $A$-invariants in these modules. \ssbegin{4.1}{Lemma} Let $\fgl(V)=\fg=\fg_-\oplus \fg_0\oplus\fg_+$, as in \S 3 and let $M$ be a $\fg_{0}$-module. Set $\fg_+M=0$. There is an isomorphism of $\fosp(V)$-modules $$ \ind^{\fgl(V)}_{\fg_0\oplus \fg_+}(M)\simeq \ind^{\fosp(V)}_{\fosp(V)_\ev}(M).\eqno{(4.1)} $$ \end{Lemma} \begin{proof} Cf. \cite{S2}, Lemma 5.1. \end{proof} \ssbegin{4.2}{Lemma} Let $\fg$ be a Lie superalgebra and the representation of $\fg_\ev$ in the maximal exterior power of $\fg_\od$ is trivial. Then there is an isomorphism of vector spaces $$ \ind^{\fg}_{\fg_\ev}(M)^{\fg}\simeq M^{\fg_\ev}.\eqno{(4.2)} $$ \end{Lemma} \begin{proof} Cf. \cite{S2}, Lemma 5.2. \end{proof} \begin{Remark} Statements similar to Lemmas 4.1 and 4.2 hold also for $U(\fosp(V))[\eps]$-modules. One can refine Lemma 4.2 and prove that if $v_0\in M$ is $\fg_\ev$-invariant, then the corresponding to it $\fg$-invariant vector is of the form $\xi_1\dots\xi_nv_0+$ terms of lesser degree. \end{Remark} \ssec{4.3} The presence of an even $\fosp(V)$- and $A$-invariant form on $V$ determins an isomorphism of $A$-modules and algebras $\fA^{p, q}_{k, l}=\fA^{p+k, q+l}$. Therefore, we may assume that $k=l=0$. By definition, the Lie superalgebra $\fosp(V)$ preserves the vector $$ \mathop{\sum}\limits_{i=1}^ne_i^*\otimes e_{n-i+1}^*+\mathop{\sum}\limits_{j=1}^r(e_{\overline{m-j+1}}^*\otimes e_{\bar j}^*-e_{\bar j}^*\otimes e_{\overline{m-j+1}}^*), $$ where $\dim V=(n|2r)$. Therefore, the scalar products $$ (v_s,v_t)=\mathop{\sum}\limits_{i=1}^nx_{is}^*x_{n-i+1, t}^*+(-1)^{p(s)}\mathop{\sum}\limits_{j=1}^r(x_{\overline{m-j+1}, s}^*x_{\bar j, t}^*-x_{\bar j, s}^*x_{\overline{m-j+1, t}}^*), \eqno{(4.3)} $$ where $s, t\in R_W$, are $\fosp(V)$- and $A$-invariants. \begin{Theorem} The algebra of $A$-invariant elements in $\fA^{p, q}=S^{\bcdot}(V^*\otimes W)$ is generated by the elements $(v_s, v_t)$ for $s, t\in R_W$. \end{Theorem} \begin{proof} Cf. \cite{S2}, Theorem 5.3. \end{proof} \ssec{4.4} Let $I$ be a sequence of length $2k$ with elements from $R_W$. Determine an element $X(I)\in S^{\bcdot}(S^2(W))$ by setting $$ X(I)=x_{i_{1}i_{2}}\dots x_{i_{2k-1}i_{2k}}, $$ where $x_{ij}$ is the canonical image of the element $w_i\otimes w_j\in S^2(W)$. Let $t$ be a tableau of order $2k$ with rows of even lengths. Then an ``even Pfaffian" is defined: $$ Pf_t(I)=\mathop{\sum}\limits_{\tau\in C_{t}, \; \sigma \in R_{t}}\eps(\tau)c(I, (\sigma \tau)^{-1})X(\sigma \tau I).\eqno{(4.4)} $$ \begin{Theorem} {\em a)} $S^{\bcdot}(S^2(W))=\oplus W^{\lambda}$, where the length of each row of $\lambda$ is even. {\em b)} Let $t$ be a $\lambda$-tableau filled in along rows with the numbers $1$, $2$, \dots . Then the family $Pf_t(I)$ for the $t$-standard sequences $I$ is a basis of $W^{\lambda}$. \end{Theorem} \begin{proof} On $T^{2k}(W)=W^{\otimes 2k}$ the group $\fS_{2k}$ and its subgroup $G_k=\fS_{k}\circ \Zee_2^ k$ naturally act; namely, $\fS_{k}$ permutes pairs $(2i-1, 2i)$ whereas $\Zee_2^ k$ permutes inside each pair. Clearly, $S^k(S^2(W))=T^{2k}(W)^{G_{k}}$. But, on the other hand, $T^{2k}(W)=\oplus\; S^\lambda\otimes W^\lambda$, so $T^{2k}(W)^{G_{k}}=\oplus\; (S^\lambda)^{G_{k}}\otimes W^\lambda$. Hence, in the decomposition of $S^k(S^2(W))$ only enter $W^\lambda$ for which $(S^\lambda)^{G_{k}}\neq 0$ and their multiplicity is equal to $\dim (S^\lambda)^{G_{k}}$. But $$ (S^\lambda)^{G_{k}}=\Hom_{G_{k}}(\ind^{\fS_{k}}_{G_{k}}(\id), S^\lambda), $$ so the multiplicity of $W^\lambda$ in $S^k(S^2(W))$ is equal to that of $S^\lambda$ in $\ind^{\fS_{k}}_{G_{k}}(\id)$. By \cite{H} it is equal to 1 if the lengths of all rows of $\lambda$ are even and 0 otherwise. This proves a). b) Consider now the natural map $T^{2k}(W)\tto S^k(S^2(W))$. For the tableau $t$ from the conditions of the theorem and the sequence $I$ the vectors $e_t(w_I)$ form a basis of $W^\lambda$. So the images of these vectors (which are exactly the $Pf_t(I)$) form a basis of $W^\lambda\subset S^k(S^2(W))$. \end{proof} \ssec{4.5} Consider the algebra homomorphism $$ S^{\bcdot}(S^2*W))\tto S^{\bcdot}(V^*\otimes W),\quad x_{st}\mapsto (v_s, v_t).\eqno{(4.5)} $$ Its kernel is the ideal of relations between scalar products. \begin{Theorem} The ideal of relations between scalar products is generated by polynomials $Pf_t(I)$, where $t$ is a $(2r+2)\times (n+1)$ rectangle filled in along rows and $I$ is a $t$-standard sequence with elements from $R_W$. \end{Theorem} \begin{proof} By Theorem 4.3 $$ S(V^*\otimes W)^A= \mathop{\oplus}\limits_\lambda(V^*{}^\lambda)^A\otimes W^\lambda= \mathop{\oplus}\limits_{\lambda_{n+1}\leq 2r}W^\lambda. $$ Hence, the kernel of homomorphism (4.5) is equal to $\mathop{\oplus}\limits_{\lambda_{n+1}\geq 2r+2}W^\lambda$. Let us prove that it is contained in the ideal generated by $W^\lambda$, where $\lambda$ is a $(2r+2)\times (n+1)$ rectangle. Let $\mu\supset\lambda$ and $e_s$ the corresponding idempotent. Then $e_s=\sum\tau_ie_t\sigma_j$. Hence, $$ e_s(J)=\mathop{\sum}\limits_{}\tau_ie_t(\sigma_jJ)= \mathop{\sum}\limits_{}e_{\tau_{i}t}(\tau_{i}\sigma_jJ). $$ Thus, $Pf_s(J)=\mathop{\sum}\limits_{i,j}f_{ij}Pf_{\tau_{i}t}(J_{ij})$ and we are done. \end{proof} \section*{\S 5. The relative invariants of \protect $\fosp(V)$} The invariants of $\fosp(V)$ are, first of all, the ones generated by scalar products. To describe the other invariants, let us describe a certain invariant in the tensor algebra. Let $\dim V= n|m$. For $i\in R_V$ define $\tilde i$ by setting $$ \tilde i= n-i+1 \text{ if $i$ is ``even" and $\overline{m-i+1}$ if $i$ is ``odd"}. $$ Let $I=i_1i_2\dots i_{2p}$ be a sequence of even length with elements from $R_V$ and $I^*$ the set consisting of the pairs $(i_{2\alpha-1}, i_{2\alpha})$ for $\alpha\leq p$ such that $\tilde i_{2\alpha-1}\neq i_{2\alpha}$. Let $t$ be a rectangular $n\times m$ tableau consequtively filled in along columns from left to right and $I$ a sequence with elements from $R_V$. Let us fill in the tableau $t$ with elements from $I$: replace $\alpha$ with $i_\alpha$. Let $\cT$ be the set of sequences $I$ such that all the rows of $t$ except the last row are of the form $$ i_1\tilde i_1\dots i_{r}\tilde i_r\text{ for }r=\frac 12 m, $$ while the last row $J$ should be such that if $j\in\hat J$, then $\tilde j\in\hat J$ and $\hat J$ consist of pairwise distinct ``odd'' elements. Let $I\in\cT$. Set $r=\frac12m$, let $\nu$ be the total amount of marked pairs from the last row consisting of pairwise conjugate ``odd" elements that do not belong to $N(L)$; let $n_1$, \dots , $n_\nu$ the multiplicities with which these pairs enter the last row and $N=n_1+ \dots+n_\nu$; let $\sigma_l$ be the $l$-th elementary symmetric function. Set $$ K(I)=\mathop{\sum}\limits_{q=s}^{s+\nu}(N+1)^r2^{r-q}(r-q)! N^q\sigma_{q-s}(n_1, \dots , n_\nu) $$ and $d(I)=d(I_1)d(I_3)\dots d(I_{2r-1})$, where $d(J)=(-1)^{\alpha(J, J)}$, see $(1.2)$??. \ssbegin{5.1}{Theorem} In $V^{\otimes n(m+1)}$ lies an $\fosp(V)$-invariant element $$ \nabla_{m+1}=\mathop{\sum}\limits_{I\in \cT}d(I)K(I)e_s(v_{I_{1}}\otimes v_{I}).\eqno{(5.1.1)} $$ \end{Theorem} \begin{proof} Set $$ c(i, \tilde i)=\cases 1&\text{ if $p(i)=0$ or $i<\tilde i$ and $p(i)=1$}\cr - 1&\text{ if $i>\tilde i$ and $p(i)=1$}.\endcases $$ The map $$ V\tto V^*, \quad e_i\mapsto c(i, \tilde i)e^*_{\tilde i} $$ is an isomorphism induced by the invariant bilinear form and $\tilde \theta_2=\mathop{\sum}\limits_{i\in R_V} c(i, \tilde i)e_i\otimes e_{\tilde i}$ is an $\fosp(V)$-invariant. Let $t$ be a rectangular $(n+1)\times m$ tableau as in Theorem 3.8 and $J$ a $t$-sequence such that after being filled each row $J$ is of the form $j_1\tilde j_1\dots j_r\tilde j_r$. Denote by $\cT_1$ the set of such sequences $J$. Then $$ \tilde \theta=\theta_2^{\otimes \frac12 (n+1)m}=\mathop{\sum}\limits_{J\in\cT_1} d(J)c(J)v_J,\eqno{(5.1.2)} $$ where $J_1$, \dots , $J_{2r-1}$ are the columns of the tableau $t$ and where $$ d(J)=d(J_1)d(J_3)\dots d(J_{2r-1})\text{ while }c(J)=c(J_1)c(J_3)\dots c(J_{2r-1}) $$ whereas $c(J_\alpha)=\mathop{\prod}\limits_{i\in J_{\alpha}} c(i, \tilde i)$. The element (5.1.2) is an $\fosp(V)$-invariant; having applied to it the operator $\cL$ from Theorem 3.8 we get another $\fosp(V)$-invariant: $$ \cL(\tilde \theta)=\mathop{\sum}\limits_{J} d(J)c(J)\cL(v_J)= \mathop{\sum}\limits_{J, l} d(J)c(J)\eps(J)\eps(J, l)e_s(v_{I_{1}}\otimes v_{J\setminus l}),\eqno{(5.1.3)} $$ where $$ \eps (J)=\mathop{\prod}\limits_{1\leq i\leq r}\eps(J_{2i-1}*J_{2i}),\quad \eps (J, l)=\sign (l)\mathop{\prod}\limits_{1\leq i\leq r}\eps(J_{2i-1}*J_{2i}, l_{2i-1}*l_{2i}). $$ For the collection $(J_{2i-1}, J_{2i}, l_{2i-1}, l_{2i})$ define the sequence $(I_{2i-1}, I_{2i})$ as follows: if $ l_{2i-1}$ and $l_{2i}$ lie in the same row just strike them out, if $ l_{2i-1}$ and $l_{2i}$ lie in distinct rows we strike them out and place their conjugates, $\tilde l_{2i-1}$ and $\tilde l_{2i}$, in the last row in the same columns. The sequence $I$ takes the form $(I_1, I_2, \dots , I_{2r-1}, I_{2r})$. It is not difficult to verify that $$ e_s(v_{I_{1}}\otimes v_{J\setminus l})=\sign(l)\eps(J, l)e_s(v_{I_{1}}\otimes v_{I})\text{ and }d(J)=(-1)^rd(I)\eps(J). $$ Therefore, $$ \cL(\tilde \theta)=(-1)^r\mathop{\sum}\limits_{J, l}\sign(l)c(J)d(I)e_s(v_{I_{1}}\otimes v_{I}). $$ The constant factor, the sign, can be, clearly, replaced with a 1. If $I$ is of the above form, then in the last row for some values of $i$ the pairs $(l_{2i-1}, l_{2i})$ are conjugate whereas all the remaining values of $i$ are odd and pairwise distinct, call them $I^*=\{k_1, \dots , k_{2p}\}$. Set $$ \hat c(I)=\sign(k_1, \dots , k_{2p})(-1)^{p}\mathop{\prod}\limits_{c(i_{2\alpha -1}, i_{2\alpha }),\neq 0}c(i_{2\alpha -1}, i_{2\alpha}), $$ where $\sign(k_1, \dots , k_{2p})$ is the sign of the permutation. Then $c(J)\eps(l)=\hat c(I)$ and, therefore, $$ \cL(\theta_{m+1})=\mathop{\sum}\limits_{J, l}c(I)d(I)e_s(v_{I_{1}}\otimes v_{I}). $$ To complete the proof, it suffices to calculate the number of pairs$(J, l)$ that give the sequence $I$ which leads to formula (5.1.1). \end{proof} \ssbegin{5.2}{Theorem} The algebra of $\fosp(V)$-invariants is generated by the polynomials {\em i)} $(v_s, v_t)$ for $s, t\in R_W$ and {\em ii)} $R(J)=\mathop{\sum}\limits_{I}d(I)K(I)Pf_s(I_1*I, J)$ for every $I\in \cT$ and every $s$-standard sequence $J$ with elements from $R_W$. \end{Theorem} \begin{proof} Let $f$ be an $\fosp(V)$-invariant which is not $A$-invariant. Let $f$ depend on $n-1$ even and $2r$ odd generic vectors $v_1$, \dots , $v_{\overline{2r}}$. Then there exists a $g\in\OSp(V\otimes A)$ such that $g\Span(v_1, \dots , v_{\overline{2r}})=\Span (e_1, \dots , e_{\overline{2r}})$. Let $he_n=-e_n$ whereas $he_i=e_i$ for $i\neq n$. Then $\ber (h)=-1$ and $f(hg\cL)=-f(g\cL)$. But, on the other hand, $f(hg\cL)=f(g\cL)$, hence, $f=0$. This means that $\fosp(V)$-invariants other than scalar products may only be of type $\lambda$ corresponding to a typical module. So, in the same vein as for $A$-invariants, we see that $\dim (V^*{}^\lambda)^{\fosp(V)}=1$ if $\lambda$ is typical and its first $n$ rows are of odd lengths whereas the remaining rows are of even lengths. If we do not consider the scalar products, then for the other (atypical) $\lambda$ there are no invariants. Under the canonical homomorphism $T^k(V^*)\otimes T^k(W)\tto S^k(V^*\otimes W)$ the module $V^*{}^\lambda\otimes W^\lambda$ turns into its copy and a basis of the first copy becomes a basis of the second one. This shows that if $\lambda$ is an $n\times (2r+1)$ rectangle, then the polynomials $R(J)$ from the theorem constitute a basis of $V^*{}^\lambda\otimes W^\lambda$, a subspace of $S^k(V^*{}\otimes W)$. For an arbitrary $\lambda$ containing an $n\times (2r+1)$ rectangle we apply the same arguments as in the proof of Theorem 2.2. \end{proof} \section*{\S 6. The invariants of \protect $\fpe(V)$} Suppose $\dim V=(n|n)$, the $e_i^*$ is a basis of $V_\ev$ and the $e_{\bar i}^*$ be the dual basis of $V_\od$ with respect to an odd nondegenerate form on $V$. Then $\fpe(V)$ preserves the tensor $\sum (e_i^*\otimes e_{\bar i}^*+e_{\bar i}^*\otimes e_i^*)$. Observe that the scalar products $$ (v_s, v_t)=\mathop{\sum}\limits_{}(-1)^{p(s)} (x_{is}^*\otimes e_{\bar i, t}^*+e_{\bar i, s}^*\otimes e_{it}^*) \text{ for any } s, t\in S $$ are $\fpe(V)$-invariants. Moreover, the presence of the odd form determines an isomorphism of algebras and $\fpe(V)$-modules $\fA^{p, q}_{k, l}=\fA^{p+l, q+k}$, so, as for the orthosymplectic case, we may assume that $k=l=0$. The compatible $\Zee$-grading of $\fgl(V)$ induces compatible $\Zee$-gradings of $\fpe(V)$ and $\fspe(V)$: $$ \fg=\fg_-\oplus\fg_0\oplus \fg_+,\text{ where $\fg_-=\Lambda^2(V)$, $\fg_+=S^2(V^*)$ and $\fg_0=\fgl(V)$ or $\fsl(V)$} $$ (There is also another, isomorphic, representation which we will not use in this paper: $$ \fg=\fg_-\oplus\fg_0\oplus \fg_+,\text{ where $\fg_-=\Lambda^2(V^*)$, $\fg_+=S^2(V)$ and $\fg_0=\fgl(V)$ or $\fsl(V)$}.) $$ Let $X_\alpha$, $1\leq \alpha\leq \frac12n(n+1)$, be a basis of $\fg_+$ and let $Y_\beta$, $1\leq \beta\leq \frac12n(n-1)$, be a basis of $\fg_-$. Set $$ X_+=\mathop{\prod}\limits_{1\leq \alpha\leq \frac12n(n+1)}X_\alpha, \quad Y_-=\mathop{\prod}\limits_{1\leq \beta\leq \frac12n(n-1)}Y_\beta. $$ Observe that the weight of $X^+$ with respect to the Cartan subalgebra is equal to $(n+1)\sum\eps_i$ and the weight of $Y^-$ is equal to $-(n-1)\sum\eps_i$. \ssbegin{6.1}{Lemma} Let $L=\ind^\fg_{\fg_0\oplus\fg_+}(M)= \ind^\fg_{\fg_0\oplus\fg_-}(N)$ be a typical irreducible $\fg=\fgl(V)$-module. There exists an isomorphism of vector spaces $$ L^{\fspe(V)}=M^{\fspe(V)_{\ev}}=N^{\fspe(V)_{\ev}} $$ given by the formulas $$ M\tto L, \; m\mapsto Y^-m\text{ and }N\tto L, \; n\mapsto X^+n $$ the inverse map being given by the formulas $$ L\tto M, \; l\mapsto X^-l\text{ and }L\tto N, \; l\mapsto Y^+l. $$ \end{Lemma} \begin{proof} Consider the two gradings of $L$: $$ L^+_k=\Span(f(X_\alpha)n: n\in N \text{ and } \deg f=k) $$ and $$ L^-_k=\Span(f(Y_\beta)m: m\in M\text{ and } \deg f=k). $$ It is clear that $L^+_k=L^-_{n^{2}-k}$. If $l$ is a $\fspe(V)$-invariant, then $X_\alpha l=0$ (for $1\leq \alpha\leq \frac12n(n+1)$) and $l=X^+f(X_\alpha)n$ for $n\in N$. Therefore, $l=\mathop{\sum}\limits_{r\geq \frac12n(n+1)}l_r^+$, where $l_r^+\in L_r^+$. We similarly establish that $l=\mathop{\sum}\limits_{1\leq s\leq \frac12n(n-1)}l_s^-$, where $l_s^-\in L_s^-$. Hence, $\mathop{\sum}\limits_{r\geq \frac12n(n+1)}l_r^+=\mathop{\sum}\limits_{1\leq s\leq \frac12n(n-1)}l_s^-$. Taking into account the equality $L^+_k=L^-_{n^{2}-k}$ we deduce that $l\in L^+_{\frac12n(n+1)}=L^-_{\frac12n(n-1)}$ and $l=X^+n=Y^-m$ for some $n\in N$ and $m\in M$. Moreover, it is clear that $m$ and $n$ are $\fspe(V)_\ev=\fsl(V_\ev)$-invariants. Conversely, if $m$ and $n$ are $\fsl(V_\ev)$-invariants, then a direct check shows that $X^+n$ and $Y^-m$ are $\fspe(V)$-invariants. \end{proof} \ssbegin{6.2}{Theorem} The algebra of $\fpe(V)$-invariants is generated by the scalar products $(v_s, v_t)$ for $s, t\in R_W$. \end{Theorem} \begin{proof} See\cite{S2}, sec. 6.2. \end{proof} Let $I$ be a sequence of length $2k$ composed of elements from $R_W$. Determine the element $Y(I)\in E^{\bcdot}(S^2(W))=S^{\bcdot}(\Pi(S^2(W)))$ by setting $$ Y(I)=(-1)^{\beta}y_{i_{1}i_{2}}\dots y_{i_{2k-1}i_{2k}}, $$ where $y_{ij}$ is the canonical image of the element $\omega_i\otimes \omega _j$ and $\beta=\mathop{\sum}\limits_{1\leq \alpha\leq k}(k-\alpha)(i_{2\alpha-1}+i_{2\alpha})$. \ssec{6.3} Let $\lambda$ be a partition of the form $(\alpha_{1}, \dots , \alpha_{p}, \alpha_{1}-1, \dots, \alpha_{p}-1)$ in Frobenius' notations (see \cite{M}) and $t$ be a tableau of the form $\lambda$ filled in so that the underdiagonal columns the diagonal cells including are filled in consequtively with odd (ili ``odd"?) numbers while the rows to the right of the diagonal are consequtively occupied by even numbers. For a tableau of such a form and a sequence $I$ the ``periplectic" Pfaffian is defined: $$ PPf_t(I)=\mathop{\sum}\limits_{\tau\in C_t, \;\sigma\in R_t}\eps(\tau)c(I, (\sigma\tau)^{-1})Y(\sigma\tau I).\eqno{(6.3.1)} $$ \ssbegin{6.3.1}{Theorem} For the above tableau $t$ the family $PPf_t(I)$ for the $t$-standard sequences $I$ is a basis in the module $W^\lambda\subset E^{\bcdot} (S^2(W))$. \end{Theorem} \begin{proof} From the theory of $\lambda$-rings it follows that $E^{\bcdot} (S^2(W))=\oplus W^\lambda$, where the sum runs over the $\lambda$ of the above described form. One can easily verify that for the tableau as indicated in the formulation of the theorem and a $t$-standard sequence $I$ the image $e_t(w_I)$ in $E^{\bcdot} (S^2(W))$ is nonzero. Hence, for a fixed tableau $t$ the canonical map $T^{2k}(W)\tto E^k(S^2(W))$ performes an isomorphism of $e_t(T^{2k}(W))$ with $W^\lambda\subset E^k(S^2(W))$. This implies the theorem. \end{proof} \ssec{6.3.2} Consider now an algebra homomorphism $$ E^{\bcdot} (S^2(W))\tto S^{\bcdot} (V^*\otimes W),\quad y_{st}\mapsto (v_s, v_t). \eqno{(6.3.2)} $$ \begin{Theorem} The kernel of $(6.3.2)$ is generated by polynomials $PPf_t(I)$, where $t$ is of the form of a $(n+1)\times(n+2)$ rectangle and is filled in as described in the previous section and $I$ is a $t$-standard sequence with elements from $R_W$. \end{Theorem} \begin{proof} Clearly, $$ (S^k(V^*\otimes W))^{\fpe(V)}= (\mathop{\oplus}\limits_{\lambda: \lambda _{n+1}\leq n}V^*{}^{\lambda}\otimes W^*{}^{\lambda})^{\fpe(V)}= \mathop{\oplus}\limits_{\lambda: \lambda _{n+1}\leq n}W^*{}^{\lambda}, $$ where $\lambda$ is of the same form as stated in Theorem. Since (6.3.2) is a $\fgl(V)$-module homomorphism, its kernel is $\mathop{\oplus}\limits_{\lambda: \lambda _{n+1}\geq n+2}W^*{}^{\lambda}$. The fact that this kernel is generated by the elements of the least degreee is proved by the same arguments as for $\fosp(V)$. \end{proof} \section*{\S 7. The invariants of $\fspe(V)$} First, let us describe certain tensor invariants. Let $\cT_1$ be the set of matrices $A$ whose entries are equal to either 1 or 0, with zeroes on the main diagonal and such that $a_{ij}+a_{ji}=1$ for all offdiagonal entries. Set $$ A_i=\sum a_{pq},\text{ where the sum runs over all the elements strictly below the $i$-th row}. $$ Define $|A|$ recursively: for $n=2$ set $|A|=0$ and for $n>2$ set $$ |A|=|A^*|+\mathop{\sum}\limits_{i=1}^{n-2}a_{in}A^*_i+ \mathop{\sum}\limits_{1\leq j<i<n}a_{in}a_{nj}+\mathop{\sum}\limits_{i>j}a_{ij}+\frac16n(n-1)(n-2), $$ where $A^*$ is obtained from $A$ by striking out the last row and the last column. \ssbegin{7.1}{Lemma} $Y^-=\mathop{\prod}\limits_{i<j}(E_{\bar i, j} - E_{\bar j, i})=\mathop{\sum}\limits_{A\in\cT_1}(-1)^{|A|}E_A$, where the product runs over the lexicografically ordered set of pairs $i<j$ and $E_A=\prod E_{\bar i, j}^{a_{i, j}}$ and where the last product is taken over the rows of the matrix $A$ from left to right and downwards. \end{Lemma} \begin{proof} Clearly, $Y^-$ is the product of $\frac 12n(n-1)$ factors. In each factor, select either $E_{\bar i, j}$ or $E_{\bar j, i}$. In the first case, for $E_{\bar i, j}$, set $a_{i, j}=1$ and $a_{j, i}=0$ in the second case set the other way round. We get a matrix with the properties desired. The sign is obtained after reordering of the sequence of the $a_{ij}$: $$ a_{12}a_{21}a_{13}a_{31}\dots a_{1n}a_{n1}\dots a_{n-1,n}a_{n, n-1}\mapsto a_{12}a_{13}\dots a_{1n}\dots a_{n-1,n}a_{n1}a_{n2}\dots a_{n, n-1}. $$ This is performed by induction: first, the pairs $a_{in}a_{ni}$ are moved to the end in increasing order, this accrues the exponent of the sign with $\frac16n(n-1)(n-2)$, then we reorder the elements with indices lesser than $n$, which adds $|A^*|$, then the elements of the sequence $a_{1n}a_{n1}a_{2n}a_{n2}\dots a_{n-1,n}a_{n, n-1}$ are rearranged into the sequence $a_{1n}\dots a_{n-1,n}a_{n1}a_{n2}\dots a_{n, n-1}$ which adds $\mathop{\sum}\limits_{j<i}a_{in}a_{ni}$ to the exponent, and, finally, the elements $a_{1n}$, \dots , $a_{n-1,n}$ are placed onto the end of the $i$-th row adding $\mathop{\sum}\limits_{i=1}^{n-2}a_{in}A^*_{i}$. Besides, if $i>j$, then $E_{\bar i, j}$ enters $Y^-$ with a minus sign; this adds $\mathop{\sum}\limits_{i>j}a_{ij}$. \end{proof} \ssec{7.2} the numbers $i$ and $j$ we be referred to as {\it conjugate} if $i=\bar j$, i.e., they are equal but belong to copies of $\Nee$ of distinct ``parity''. Let $\cT_2$ be the set of sequences of length $n^2$ considered as $n\times n$-tableaux filled in along columns and with the following properties: the numbers symmetric with respect to the main diagonal are conjugate, the $(i, j)$-th entry is occupied with one of the numbers $i$ or $\bar j$, the main diagonal is filled in with ``odd" numbers $\bar 1$, \dots , $\bar n$. For every $L\in\cT_2$ determine the matrix $A=(a_{ij})$ by setting $a_{ij}=p(l_{ij})$, $n(L)=\#(\text{``even" elements in }L)$, $m(L)=m(A)=\mathop{\sum}\limits_{i\text{ is ``even"}}a_{ij}$, $\eps(L)=(-1)^{|A|+n(L)}$; let $m_k(L)=\frac{((n+k)!)^n}{(n+k-l_1)!\dots (n+k-l_n)!}$, where $l_i=\sum_j a_{ij}$. \begin{Theorem} The elements $$ e_t\left (\mathop{\sum}\limits_{}(-1)^{km(L)}\eps(L)m_k(L)v^*_L\otimes v^*_{J_{k}}\right)\text{ for }L\in\cT_2 $$ and $$ e_t\left (\mathop{\sum}\limits_{}\eps(L)m_0(L)v^*_L\otimes v^*_{I_{k}}\right )\text{ for }L\in\cT_2 $$ are $\fspe(V)$-invariant. \end{Theorem} \begin{proof} Let $r$ be an $(n+k)\times n$ rectangle filled in along columns. Set $w=v^*_{J_{n+k}}$. Denote by $w^{j_{1}}_{i_{1}}\dots w^{j_{p}}_{i_{p}}$ the tensor obtained from $w$ by replacing the elements occupying positions $i_{1}$, \dots , $i_{p}$ with numbers $j_{1}$, \dots , $j_{p}$, respectively. Then $$ e_r(E_{\bar n, j}w)=(-1)^{i-1}e_r(w^{j}_{i})(n+k), $$ where $i$ is any of the numbers of the positions occupied by $\bar n$. If $E_{A_{n}}=\prod E_{\bar n, j}^{\alpha_{n, j}}$, the product being ordered in order of increase of indices $j$, then $$ e_r(E_{A_{n}}w)=(-1)^{i_{1}-1+ \dots + i_{l}-1}\frac{(n+k)!}{(n+k-l)!}e_r(w^{j_{1}, \dots , j_{l}}_{i_{1}, \dots , i_{l}}), $$ where $\{j_{1}, \dots , j_{l}\}=\{j\mid \alpha_{n, j}\neq 0\}$ and $l=\mathop{\sum}\limits_{j}\alpha_{n, j}$, where $i_{1}< \dots < i_{l}$. Assume that $\{i_{1}, \dots , i_{l}\}=\{a+j_{1}-1, \dots , a+j_{l}-1\}$, where $a$ is the number of the first element in the $n$-th column of tableau $r$. We thus get $$ e_r(E_{A_{n}}w)=(-1)^{l\cdot a+ j_{1}+\dots + j_{l}}\frac{(n+k)!}{(n+k-l)!}e_r(w^{j_{1}, \dots , j_{l}}_{i_{1}, \dots , i_{l}}). $$ By continuing the process we get $$ e_r(E_{A}w)=(-1)^{\eps(A)}\frac{[(n+k)!]^n}{(n+k-l_{1})!\dots (n+k-l_{n})!}e_r(v_{I_{A}}^*), $$ where $\eps(A)=a_1+\dots +a_n+n(A_{n})+n(A_{n-1})+\dots +n(A_{1})$ and $a_{i}$ is the number of the first element in the $i$th column and where $n(A_{i})$ is equal to the sum of the numbers of the places occupied by the 1's, and where $I_{A}$ coinsides with $J_{n+k}$ everywhere unless the $a_{ij}=1$; then the $(ij)$-th entry of $I_{A}$ is occupied by $j$. If $I$ and $J$ are two sequences and $t$ and $s$ are two tableaux of the same form such that after filling $t$ with the elements from $I$ and $s$ with the elements from $J$ one gets geometrically identical pictures, then $\sigma t=s$ implies $\sigma I=J$. Indeed, $$ I(\sigma^{-1}\alpha)=t(\sigma^{-1}\alpha)=s(\alpha)=J(\alpha). $$ Therefore, if $\sigma t=r$, we have $$ \renewcommand{\arraystretch}{1.4} \begin{array}{l} Y^-e_t(v^*_{J_{n}}v^*_{J_{k}})= Y^-e_{\sigma^{-1}r}(v^*_{\sigma^{-1}(J_{n+k})})=\\ Y^-\sigma^{-1}e_{r\sigma}(\sigma^{-1}v^*_{J_{n+k}})\cdot c(J_{n+k}, \sigma)= c(J_{n+k}, \sigma)\sigma^{-1}Y^-e_{r}(v^*_{J_{n+k}}) \end{array} $$ because thanks to the fact that $J_{n+k}$ only contains ``odd" elements $c(J_{n+k}, \sigma)=\sign (\sigma)$. Hence, $$ \renewcommand{\arraystretch}{1.4} \begin{array}{l} Y^-e_t(v^*_{J_{n}}v^*_{J_{k}})= \sign(\sigma)\sigma^{-1}e_r\left (\mathop{\sum}\limits_A(-1)^{\eps(A)} \frac{[(n+k)!]^n}{(n+k-l_{1})!\dots (n+k-l_{n})!} v_{I_{A}}^*\right )=\\ \sign(\sigma)e_t\left (\mathop{\sum}\limits_A(-1)^{\eps(A)} \frac{[(n+k)!]^n}{(n+k-l_{1})!\dots (n+k-l_{n})!} \sigma^{-1}v^*_{I_{A}}\right )= \\ \sign(\sigma)e_t\left (\mathop{\sum}\limits_A(-1)^{\eps(A)} \frac{[(n+k)!]^n}{(n+k-l_{1})!\dots (n+k-l_{n})!} c(I_{A}, \sigma)v^*_{J_{A}}\otimes v_{J_{k}}\right ). \end{array} $$ where $$ c(I_{A}, \sigma)=|A_{2}|\cdot k+|A_{4}|\cdot k\dots = k\mathop{\sum}\limits_{i \text{ is even}}a_{ij}, $$ and where $J_{A}$ coinsides with $J_{n}$ everywhere unless where $a_{ij}=1$, then the $(i,j)$the position is occupied by $j$. We are done. \end{proof} \ssbegin{7.3}{Theorem} The algebra of $\fspe(V)$-invariant polynomials is generated by the following elements {\em i)} $(v_\alpha, v_\beta)$ for $\alpha, \beta\in R_W$; {\em ii)} $PPf_{k}(J)=\mathop{\sum}\limits_{} (-1)^{(k-1)m(L)}\eps(L)m_{k-1}(L)P_t(L*J_k, J)$ for any $t$-standard sequence $J$; {\em iii)} $PPf_{-k}(J)=\mathop{\sum}\limits_{}\eps(L)m_{0}(L)P_t(L*I_{k+1}, J)$ for any $s$-standard sequence $J$, \noindent where $k\geq 1$ and sums run over $l\in\cT_2$. \end{Theorem} \begin{proof} is similar to that of 5.2. \end{proof}
\section*{Introduction} When J.H.C. Whitehead wrote his famous papers on ``Combinatorial Homotopy'', \cite{jh1}, it would seem that his aim was to produce a combinatorial, and thus potentially constructive and computational, approach to homotopy theory, analogous to the combinatorial group theory developed earlier by Reidemeister and others. In those papers, he introduced CW-complexes and also the algebraic `gadgets' he called homotopy systems, and which are now more often called free crossed complexes, \cite{bh1}, or totally free crossed chain complexes, \cite{baues1}. Another algebraic model for a (connected) homotopy type is a simplicial group and again, there, one finds a notion of freeness. In both cases we have `freeness', yet no easily defined category of things on which our objects are `free'. Kan, \cite{kan2}, introduced the notion of a CW-basis for a free simplicial group and more recently, \cite{rab}, R.A.Brown has introduced Peiffer-Whitehead word systems or extended group presentations as a means of presenting a `homotopy system'. In both cases the aim is to use the `generators' as a combinatorial way of controlling or manipulating the algebraic model, i.e. extending the `yoga' of combinatorial group theory to higher dimensions. There are ways of passing from a simplicial group to a crossed complex (see for example, \cite{ep}) and as these are all equivalent to a left adjoint, one expects freeness to be preserved, and it is, but this is not trivial. As we do not know on what type of thing the simplicial group is free, nor on what the crossed complex is free, the conclusion is not a simple consequence of left adjointness of some sort. The problem is that to construct the $n$\textsuperscript{th} level, you need some generators together with a map to the $(n-1)$\textsuperscript{st} one, and of course you cannot do that until that level is constructed! In this paper we apply methods from our earlier papers \cite{mp,mp1}, to examine the relationship between the notions of free basis for simplicial groups and that for crossed complexes. We have included a shortened proof of the result from \cite{carrasco} and \cite{ep}, describing the passage from simplicial groups to crossed complexes, as this allows for a direct verification of freeness at the base of the crossed complex. Although our results would seem to apply in general, we have restricted detailed attention to simplicial resolutions. This is partially since there are known problems of non-realizability of a homotopy system by a CW-complex (cf. Whitehead, \cite{jh1}, section 15, or R.A.Brown, \cite{rab}, p.527) and hence by a free simplicial group with CW-basis. It thus seems prudent to understand these non-realizability results better from this simplicial viewpoint before attempting to look at the general case. Those results do not seem to disturb the general case in any significant way, but they leave them somewhat incomplete in the view they give of the general problem. \section{Preliminaries } We will denote the category of groups by $\mathfrak{Grp}.$ \subsection{Simplicial Groups} A simplicial group ${\bf G}$ is a simplicial object in the category of groups. We will denote the category of simplicial groups by $\mathfrak{SimpGrp}.$ We will only need a small amount of the extensive theory of simplicial groups here and would refer to the book by May \cite{may} or the survey by Curtis \cite{curtis} for information on the more `classical' parts of the theory. We will assume a basic knowledge of the elementary homotopy of simplicial sets and simplicial groups, but will also refer to facts and concepts from earlier parts of this series of papers, \cite{mp,mp1,mp2}. If ${\bf G}$ is a simplicial group, then $(NG,\partial)$ will be the corresponding Moore complex. Our conventions on this and related notions are given in \cite{mp}. \subsection{Step By Step Constructions} This section is a brief r\'esum\'e of how to construct simplicial resolutions. The work depends heavily on a variety of sources, mainly \cite{andre}, \cite{keune} and \cite{moore}. Andr\'e only treats commutative algebras in detail, but Keune \cite{keune} does discuss the general case quite clearly. First recall the following notation and terminology which will be used in the construction of a simplicial resolution. Let $[n]$ be the ordered set, $[n]=\{0<1<\cdots<n\}$. We define the following maps: Firstly the injective monotone map $\delta _i^n:[n-1]\rightarrow [n]$ is given by $$ \delta _i^n(x)=\left\{ \begin{array}{lcl} x & \text{if} & x<i, \\ x+1 & \text{if} & x\geq i, \end{array} \right. $$ for $0\leq i\leq n\neq 0.$ An increasing surjective monotone map $\alpha_{i}^{n} : [n+1] \to [n]$ is given by \[ \alpha_{i}^{n}(x) = \left\{ \begin{array}{ll} x & \mbox{ if $ x \leq i, $ } \\ x-1 & \mbox{ if $ x > i, $ } \end{array} \right. \] for $0 \leq i \leq n$. We denote by $\{m,n\}$ the set of increasing surjective maps $[m]\rightarrow [n].$\\ \medskip \textbf{Killing Elements in Homotopy Groups}\\ The following section describes the `step-by-step' construction due to Andr\'e \cite{andre}, that source however concentrates on simplicial algebras. We have adapted his treatment to handle simplicial groups. We recall that if $F$ and $G$ are groups, a map \[ G \times F \to F \] \[(g,f) \longmapsto{}^{g} f \] is a left action if and only if for all $g,{g'} \in G, f,{f'}\in G,$ $1.$ ${}^{g}(f {f'}) ={}^{g}f{~}{}^{g}{f'},$ $2.$ ${}^{g{g'}}f = {}^{{g}}({}^{g'}f),$ $3.$ ${}^{1}f = f.$ \\ In this case we say $F$ is a $G$-group. Let {\bf G} be a simplicial group and let $k\geq 1$ be fixed. Suppose we are given a set $\Omega $ of elements $ \Omega = \{x_\lambda :\lambda \in \Lambda \}$, $x_\lambda \in \pi _{k-1}({\bf G}),$ then we can choose a corresponding set of elements ${\it \vartheta}_\lambda \in NG_{k-1}$ so that $x_\lambda ={\it \vartheta}_\lambda{~}\partial _k(NG_k).$ (If $k=1,$ then as $NG_0=G_0,$ the condition that ${\it \vartheta}_\lambda \in NG_0$ is immediate.) We want to `kill' the elements in $ \Omega$. We form a new simplicial group $F_n$ where 1) $F_n$ is the free $G_n$-group, $$ F_n= \coprod_{\lambda,t}G_n\{y_{\lambda ,t}\}\ \text{with}\ \lambda \in \Lambda \text{ and }\ t\in \{n,k\}, $$ where $G_n\{y\} =G_n*<y>,$ the free product of $G_n$ and a free group generated by $y.$ 2) For $0\leq i\leq n$, the group homomorphism $s_i^n:F_n\rightarrow F_{n+1}$ is obtained from the homomorphism $s_i^n:G_n\rightarrow G_{n+1}$ with the relations \begin{center} $s_i^n(y_{\lambda ,t})=y_{\lambda ,u}$ \ \ with \ \ $u=t\alpha_i^n,\ \ t:[n]\rightarrow [k].$ \end{center} 3) For $0\leq i\leq n\neq 0,$ the group homomorphism $d_i^n:F_n\rightarrow F_{n-1}$ is obtained from $d_i^n:G_n\rightarrow G_{n-1}$ with the relations $$ d_i^n(y_{\lambda ,t})=\left\{ \begin{array}{clcl} y_{\lambda ,u} & \text{if the map} & u=t\delta _i^n & \text{is surjective}, \\ t{^{\prime }}({\it \vartheta}_\lambda ) & \text{if } & u=\delta _k^kt{^{\prime }}, & \\ 1 & \text{if } & u=\delta _j^kt{^{\prime }} & \text{with}\ \ j\neq k, \end{array} \right. $$ by extending multiplicatively. We sometimes denote the ${\bf F} $ so constructed by ${\bf G}(\Omega)$. \medskip {\bf Remark : } In a `step-by-step' construction of a simplicial resolution, (see below), there are thus the following properties: i) $F_n= G_n$ for \mbox{$n<k$,} ii) $F_k=$ a free $G_k$-group over a set of non-degenerate indeterminates, all of whose faces are the identity except the $k^{th}$, and iii) $F_n$ is a free $G_n$-group on some degenerate elements for $n>k.$ We have immediately the following result, as expected. \begin{prop}\label{veli} The inclusion of simplicial groups ${\bf G}\hookrightarrow{\bf F}$, where ${\bf F}={\bf G}(\Omega )$, induces a homomorphism $$ \pi_n({\bf G})\longrightarrow \pi_n({\bf F}) $$ for each $n$, which for $n<k-1$ is an isomorphism, $$ \pi_n({\bf G})\cong \pi_n({\bf F}) $$ and for $n=k-1$, is an epimorphism with kernel generated by elements of the form $\bar \vartheta_\lambda = \vartheta_\lambda \partial _kNG_k$, where $\Omega = \{x_{\lambda}: \lambda\in\Lambda\}.$ \end{prop}\hfill$\square$ \textbf{Constructing Simplicial Resolutions}\\ The following result is essentially due to Andr\'{e} \cite{andre}. \begin{thm} If $G$ is a group, then it has a free simplicial resolution $\mathbb{F}$. \end{thm} \begin{pf} The repetition of the above construction will give us the simplicial resolution of a group. Although `well known', we sketch the construction so as to establish some notation and terminology. Let $G$ be a group. The zero step of the construction consists of a choice of a free group F and a surjection $g : F\rightarrow G$ which gives an isomorphism $F/{\rm Ker}g\cong G$ as groups. Then we form the constant simplicial group ${\mathbf F}^{(0)}$ for which in every degree $n,$ $F_n=F$ and $d_i^n=$ id $=s_j^n$ for all $i,j.$ Thus ${\mathbf F}^{(0)}={\bf K}(F,0)$ and $\pi _0({\mathbf F}^{(0)})=F.$ Now choose a set $\Omega^0$ of normal generators of the normal subgroup $N={\rm Ker}(F\stackrel{g}{\longrightarrow}G),$ and obtain the simplicial group in which $F_1^{(1)}=F(\Omega ^0)$ and for $n>1,$ $F_n^{(1)}$ is a free $F_n$-group over the degenerate elements as above. This simplicial group will be denoted by $\bf{F}^{(1)}$ and will be called the {\em 1-skeleton of a simplicial resolution of the group} $G$. The subsequent steps depend on the choice of sets, $\Omega^0$, $\Omega^1, \Omega^2, \ldots, \Omega^k,\ldots .$ Let ${\mathbf F}^{(k)}$ be the simplicial group constructed after $k$ steps, the $k$-skeleton of the resolution. The set $\Omega^k$ is formed by elements $a$ of $F_k^{(k)}$ with $d_i^k(a)=1$ for $0\leq i\leq k$ and whose images $\bar a$ in $\pi _k({\mathbf F}^{(k)})$ generate that module over $F_k^{(k)}$ and ${\bf F}^{(k+1)}$. Finally we have inclusions of simplicial groups $$ {\mathbf F}^{(0)}\subseteq {\mathbf F}^{(1)}\subseteq\cdots \subseteq {\mathbf F}^{(k-1)}\subseteq {\mathbf F}^{(k)}\subseteq \cdots $$ and in passing to the inductive limit (colimit), we obtain an acyclic free simplicial group ${\mathbf F}$ with ${D}_n = F_n^{(k)}$ if $n\leq k.$ ${\mathbb F} = ({\mathbf F},g)$ is thus a simplicial resolution of the group $G$. The proof of theorem is completed.~\end{pf} \medskip {\bf Remark : } A variant of the `step-by-step' construction gives: if ${\bf G}$ is a simplicial group, then there exists a free simplicial group ${\mathbf F}$ and an epimorphism $ {\mathbf F}\longrightarrow {\bf G} $ which induces isomorphisms on all homotopy groups. The details are omitted as they are well known. \medskip {\bf Terminology : } It is sometimes useful to write ${\mathbb F}^{(k)} = ({\mathbf F}^{(k)},g)$ for the augmented simplicial group constructed at the $k^{th}$ step. The data needed to go from $\mathbb{F}^{(k)}$ to $\mathbb{F}^{(k+1)}$ are more precisely a set $\Omega^k$ and a function $g^{(k)}: \Omega^{k}\longrightarrow F_k^{(k)}$ whose image is contained in $NF_k^{(k)}$ and which generates $\pi_k(\mathbb{F}^{(k)}).$~ (We often consider $g^{(k)}$ as being an inclusion and leave it out of the notation.) The pair $(\Omega^k, g^{(k)})$ is then called {\em k-dimensional construction data for the resolution} and the finite sequence $ ((\Omega^0,g^{(0)}),\ldots,(\Omega^{k-1},g^{(k-1)})) $ is called a $k$\textsuperscript{th}-{\em level presentation} of the group $G$. The key observation, which follows from the universal property of the construction, is a freeness statement: \begin{prop}\label{free} Let $\mathbf{F}^{(k)}$ be a $k$-skeleton of a simplicial resolution of $G$ and $(\Omega^k, g^{(k)})$ $k$-dimension construction data for $\mathbb{F}^{(k+1)}.$ Suppose given a simplicial group morphism $\Theta:\mathbf{F}^{(k)}\longrightarrow {\bf G}$ such that $\Theta_{\ast} (g^{(k)}) =0,$ then $\Theta$ extends over $\mathbb{F}^{(k+1)}.$ \end{prop} This freeness statement does not contain a uniqueness clause. That can be achieved by choosing a lift for $\Theta_kg^{(k)}$ to $NG_{k+1},$ a lift that must exist since $\Theta_{\ast}(\pi_k(\mathbb{F}^{(k)}))$ is trivial. When handling combinatorially defined resolutions, rather than functorially defined ones, this proposition is as often as close to `left adjointness' as is possible without entering the realm of homotopical algebra to an extent greater than is desirable for us here. We have not talked here about the homotopy of simplicial group morphisms, and so will not discuss homotopy invariance of this construction for which one adapts the description given by Andr\'e, ~\cite{andre}, or Keune, \cite{keune}. \subsection{CW-bases} We recall from \cite{kan2} and \cite{curtis} the following definitions. \begin{defn} A simplicial group {\bf F} will be called free if\\ (a)\qquad $F_n$ is a free group with a given basis, for every integer $n\geq 0,$\\ (b)\qquad The bases are stable under all degeneracy operators, i.e., for every pair of integers $(i,n)$ with $0\leq i\leq n$ and every given generator $x\in F_n$ the element $s_i(x)$ is a given generator of $F_{n+1}.$ \end{defn} \begin{defn} Let ${\bf F}$ be a free simplicial group (as above). A subset $\mathfrak{F}\subset {\bf F}$ will be called a $CW-basis$ for ${\bf F}$ if \\ (a)\qquad $\mathfrak{F_n} = \mathfrak{F}\cap F_n$ freely generates $F_n$ for all $n\geq 0,$\\ (b)\qquad $\mathfrak{F}$ is closed under degeneracies, i.e., $x\in \mathfrak{F_n}$ implies $s_i(x)\in \mathfrak{F_{n+1}}$ for all $0\leq i\leq n,$\\ (c)\qquad if $x\in\mathfrak{F_n}$ is non-degenerate, then $d_i(x) = e_{n-1}$, $(e_{n-1}$, the identity element of $F_{n-1}$) for all $0\leq i< n.$ \end{defn} Let $\bf{F}$ be a free simplicial group with given $CW$-basis, $\mathfrak{F},$ then $X_0=\mathfrak{F}_0$ freely generates $F_0$, that is, $F_0 = F(X_0).$ In general, note that if $Y_n = \mathfrak{F}_n\setminus\bigcup\limits_{i=0}^{n-1}s_i(\mathfrak{F}_i)$ then $Y_n\subseteq NF_n$. \subsection{Crossed Modules} J. H. C. Whitehead $(1949)$ \cite{jh1} described crossed modules in various contexts especially in his investigation into the group structure of relative homotopy groups. \begin{defn} A {\rm pre-crossed module of groups} consists of a group, $G_1,$ a $G_1$-group $G_2$, and a group homomorphism $ \partial :G_2\longrightarrow G_1, $ such that for all $g_2 \in G_2,g_1 \in G_1$ $\\ \begin{array}{cccc} CM1)\ \ \ \ & \partial ({}^{g_1}g_2) & = & g_1\partial(g_2){g_1}^{-1}. \end{array} $\\ This is {\rm a crossed module }if in addition, for all $g_2,g_2^{\prime }\in G_2$,\\ $ \begin{array}{cccc} CM2)\ \ \ \ & {}^{\partial(g_2)}{g_2'} & = & g_2{g_2'}(g_2)^{-1}. \end{array} $ \end{defn} The second condition (CM2) is called {\em the Peiffer identity}. We denote such a crossed module by $(G_2,G_1,\partial )$. Clearly any crossed module is a pre-crossed module. \subsection{Free Crossed Modules} The notion of a free crossed module was described by J. H. C. Whitehead ~\cite{jh1}. We refer the reader to \cite{bh3} for the construction of a free crossed module on a presentation and the proofs of the results below. The related notion of totally free (pre-)crossed module is discussed in \cite{mp2}. \begin{thm}\label{rin} A free crossed module $G_1$-module $(G_2,G_1,\partial )$ exists on any function $f:S\rightarrow G_1$ with codomain $G_1.$ \end{thm} \begin{pf} See ~\cite{bh3}.~ \end{pf} \medskip If ($G_2,G_1,\partial $) is a free crossed $G_1$-module on the trivial function $$1:S\rightarrow G_1,$$ then $G_2$ is a free $G_1$-module on the set $S$. \section{Crossed Complexes} \subsection{Peiffer pairings and boundaries in the Moore complex} Firstly we recall from ~\cite{mp} the following result. Let ${\bf G}$ be a simplicial group with Moore complex ${\bf NG}$ and for $n \geq 1$, let $D_{n}$ be the normal subgroup generated by the degenerate elements of $n$. If $G_{n} \ne D_{n}$, then \[ NG_{n}\cap D_n = N_{n} \cap D_n \quad\mbox{ for all $n \geq 1,$ } \] where $N_n$ is a normal subgroup in $G_n$ generated by an explicitly given set of elements. \subsection{Crossed Complexes and Crossed Resolutions} The definition of a crossed complex (over a groupoid) was given by R. Brown and P. J. Higgins (1981)~\cite{bh1} generalising earlier work of Whitehead (1949)~\cite{jh1}. Crossed resolutions are discussed in several sources. A particularly useful one is the thesis of Tonks, \cite{tonks}, which handles constructions of crossed resolutions in some detail. \subsection{Peiffer-Whitehead word systems} R.A.Brown, \cite{rab}, introduces a system of generators for a `homotopy system' that he calls a {\em Peiffer-Whitehead word system}. His sets of generators are only in a finite number of dimensions whilst ours may need to be in an infinite set of levels to get a resolution, so his needs are not the same as ours, but nonetheless it seems worthwhile to include his definition as it provides a point of comparison with his work: \begin{defn}{\rm \cite{rab}, p. 525} A {\em Peiffer-Whitehead word system} or {\em extended group presentation } $ W$ consists of a finite list of finite sets $\langle W^{(1)}|W^{(2)}|\ldots W^{(n)}\rangle$ together with {\em boundary homomorphisms} $d_3, \ldots, d_n$ described as follows:\\ $W^{(1)} = I_1$ is a set of indices;\\ $W^{(2)} = \{w^2_\beta|\beta \in I_2\}$ is a set of words representing elements of the free group $F = F(I_2)$;\\ $W^{(3)} = \{w^3_\gamma|\gamma \in I_3\}$ is a set of words representing elements of the free $F$-crossed module $C(I_2)$ with boundaries $\{c_\beta =\langle w^2_\beta\rangle | \beta \in I_2\}$;\\ $W^{(m)} = \{w^m_\mu|\mu \in I_m\}\quad (4< m \leq n)$ is a set of words representing elements of the free $\mathbb{Z}G$-module $M_m = M_m(I_{m-1})$, where $G$ is the group presented by $\langle W^{(1)}|W^{(2)}\rangle$;\\ $d_3 : C(I_2) \rightarrow F$ is a group homomorphism determined by $d_3 (i_\beta) = \langle w^2_\beta \rangle$;\\ $d_4 : M_4(I_3) \rightarrow C(I_2)$ is a homomorphism determined by $d_4 (i_\gamma) = \langle w^3_\gamma \rangle$;\\ $d_m : M_m(I_{m-1}) \rightarrow M_{m-1} \quad (4< m \leq n)$ is a module homomorphism determined by $d_m(i_\lambda) = \langle w^{m-1}_\lambda \rangle$;\\ In addition all words must have trivial boundaries $$d_m\langle w^m_\mu \rangle = identity \mbox{ \quad for } \mu \in I_m, 3\leq m \leq n.$$ \end{defn} Such a word system clearly specifies the generators of each level and their images in the next level down. \subsection{From Simplicial Groups to Crossed Complexes}\label{fa} P. Carrasco and A. M. Cegarra ~\cite{c:c} defined $$ C_n({\bf G})=\frac{NG_n}{(NG_n\cap D_n)~d_{n+1}(NG_{n+1}\cap D_{n+1})} $$ for a simplicial group {\bf G}. This gives a crossed complex ${\mathfrak{C}}({\bf G})$ starting from the Moore complex ({\bf NG}, $ \partial )$ of ${\bf G}.$ The map $ \partial _n:C_n({\bf G})\rightarrow C_{n-1}({\bf G}) $ will be that induced by $d_n^n$. Their proof requires an understanding of hypercrossed complexes. P. J. Ehlers and T. Porter, ~\cite{ep}, developed a more direct proof for simplicial groupoids. Here we will sketch a shorter argument showing that $ {\mathfrak{C}}({\bf G})$ is a crossed complex as we can use some of the ideas later on. This proof emphasises the role played by the various $F_{\alpha,\beta}.$ These pairing operations on the Moore complex were introduced in \cite{mp} and \cite{mp1}. They are defined by forming $[s_\alpha x,s_\beta y]$ and then projecting the result into the Moore complex. Detailed examples and calculations are given in these papers cited above. If $x\in NG_n$, we will write $\bar x$ for the corresponding element of $C_n({\bf G})$. \begin{lem} The subgroup $(NG_n\cap D_n)~ d_{n+1}(NG_{n+1}\cap D_{n+1})$ is a normal subgroup in $G_n$. \end{lem} \begin{pf} This is a routine use of the degeneracies. \end{pf} \begin{prop} Let {\bf G} be a simplicial group, then defining $C({\bf G})= (C_n({\bf G}), \partial)$ as above yields a crossed complex. \end{prop} \begin{pf} (i) For $n\geq 2,$ $C_n({\bf G})$ is abelian, in fact $$ \begin{array}{rcl} F_{(n-1)(n)}(x,y)& = &[s_{n-1}x, ~s_ny]~[s_ny, ~s_nx]\\ d_{n+1}F_{(n-1)(n)}(x,y)& = & [x, ~s_{n-1}d_n(y)]~[y,x] \end{array}$$ is in $(NG_n\cap D_n)d_{n+1}(NG_{n+1}\cap D_{n+1}), $ so $d_{n+1}F_{(n-1)(n)}(x,y) \equiv 1 $ mod $(NG_n\cap D_n)d_{n+1}(NG_{n+1}\cap D_{n+1})$ giving $\overline{x}\overline{y} = \overline{y}\overline{x}.$\\ (ii) For $x\in NG_n,$ and $y\in NG_m,$ taking $\alpha=(n,n-1,\ldots,m)$, and $\beta =(m-1),$ it is easy to see that\\ $$ F_{\alpha, \beta}(x_{\alpha}, y_{\beta}) = \prod\limits_{k=0}^{n-m+1} [s_ns_{n-1}\ldots s_mx, ~s_{m-1+k}y]^{(-1)^k} [s_ns_{n-1}\ldots s_m(x),~s_{n}y]$$ $$d_{n+1}F_{\alpha, \beta}(x_{\alpha}, y_{\beta}) = \prod\limits_{k=0}^{n-m} [s_ns_{n-1}\ldots s_mx, ~s_{m-1+k}y]^{(-1)^k} [s_{n-1}\ldots s_m(x),~y]. $$ This implies that $[s_m^{(n-m)}(x), y] \in (NG_n\cap D_n)d_{n+1}(NG_{n+1}\cap D_{n+1}),$ (where $s_{m}^{(n-m)}x = s_{n-1}\ldots s_mx$) which shows that the actions of $NG_m$ on $NG_n$ defined by conjugation $$ {}^{\overline{x}}\overline{y}= \overline{s_{m}^{(n-m)}(x)ys_{m}^{(n-m)}(x)^{-1}} $$ via these degeneracies are trivial if $m\geq 1.$ For $m=1,$ this gives $\alpha =(n,n-1,\ldots,1)$, $\beta=(0)$ and\\ \hspace*{1.5cm}$ F_{(n,n-1\ldots,1)(0)}(x, y) = $\\ \hspace*{3cm}$ \prod\limits_{k=0}^{n} [s_ns_{n-1}\ldots s_1x, ~s_{0+k}y]^{(-1)^k} [s_ns_{n-1}\ldots s_1(x),~s_{n}y],$ \\where $x\in NG_1, y\in NG_n,$ and it is easily checked that\\ \hspace*{1.5cm} $ d_{n+1}F_{(n,n-1\ldots,1)(0)}(x, y) = $\\ \hspace*{3cm}$ \prod\limits_{k=0}^{n-1} [s_{n-1}\ldots s_1x, ~s_{0+k}d_ny]^{(-1)^k}[s_ns_{n-1}\ldots s_1(x),~s_{n}y]. $\\ Then $ [s_ns_{n-1}\ldots s_1(x),~s_{n}y]\equiv 1\qquad \mbox{mod}~(NG_n\cap D_n)d_{n+1}NG_{n+1}\cap D_{n+1}). $ This gives the following if $\bar{x}\in C_1$ then $\bar{x}$ and $\partial_1\bar{x}$ acts on $C_n$ in the same way and so $\partial_1C_1$ acts trivially on $C_n.$\\ (iii) This axiom follows since $$ C_1({\bf G}) = \dfrac{NG_1}{\partial_2(NG_2\cap D_2)} = \dfrac{NG_1}{[\mbox{Ker}d_1,~\mbox{Ker}d_0]} $$ and $[\mbox{Ker}d_1, \mbox{Ker}d_0]$ contains the Peiffer elements so $(C_1({\bf G}), C_0({\bf G}), \partial)$ is a crossed module. (iv) By defining $$\partial_n\bar{z} = \overline{d_n^n(z)} \quad\mbox{with} ~~z\in NG_n$$ one obtains a well defined map $\partial: C_n({\bf G})\longrightarrow C_{n-1}({\bf G})$ satisfying $\partial\partial =1.$ \end{pf} One of the immediate consequences of the above is that if ${\bf {\mathbb G}} = ({\mathbf G},f)$ is a simplicial group augmented over a group $G$, then ${\mathfrak{C}} = (C({\bf G}), f)$ is an augmented crossed complex over $G$. Moreover if ${\bf {\mathbb G}}$ is exact at $G_0$, then ${\mathfrak{C}}$ is also exact at $C_0({\bf G})$. Thus to study what happens to a resolution we need only consider the freeness and exactness in higher dimensions. \section{`Step-by-Step' Constructions and CW-bases} In this section, we describe the special case of the `step-by-step' construction of a free simplicial resolution and its skeleton up to dimension 2 and will interpret this construction and see how that relates to other constructions such as that of a free crossed module. Many of the observations that we will make, do apply more generally to arbitrary free simplicial groups with specified CW-basis, but our aim here is limited to examining resolutions in some detail. We first examine the relationship of a CW-basis to the step-by-step construction given earlier. The 1-skeleton ${\bf{ F}}^{(1)}$ of a free simplicial resolution of a group $G$ was built by adding new indeterminates, for instance, in one to one correspondence with $\Omega^{1}$ a set of generators for $\pi_1({F}^{(0)})$, $F_1^1 = F_1^{(0)}(X_0) = F(s_0(X_0) \cup Y_1) \cong F(s_0(X_0))\ast F(Y_1)$, where $\ast$ is free product, with the face maps and degeneracy map% $$ \diagram {F}(s_0(X_0)\cup Y_1)\rto<0.25ex>\rto<1ex>^{\qquad d_0,d_1 } & {F(X_0)} \lto<0.75ex>^{\qquad s_0}\rto^{\quad d_0^0}&G \enddiagram $$ where $F(X_0)\stackrel{d_{0}^{0}}{\longrightarrow} G$ is an augmentation map and $s_0,$ $d_0^1$ and $d_1^1$ are given by $$ \begin{array}{ccc} d_1^1(y_i)=b_i\in \text{Ker}d_0^0, & d_0^1(y_i)=1, & s_0(x_0)=s_0(x_0) ~~~\mbox{for}~~~ x_0\in X_0. \end{array} $$ We note that this makes $\langle X_0 \mid d_1Y \rangle$ into a presentation of $G$ in the ordinary sense. The 1-skeleton ${\bf {\bf{F}}}^{(1)}$ looks like:% $$ \diagram{...\quad F(s_1s_0(X_0)\cup s_0(Y_1)\cup s_1(Y_1)) \rto<0.25ex> \rto<1ex> \rto<1.75ex>^{\qquad\qquad d_0 ,d_1 ,d_2 } & F(s_0(X_0)\cup (Y_1)) \lto<0.75ex> \lto<1.50ex>^{\qquad\qquad s_1,s_0}\rto<0.25ex> \rto<1ex>^{\qquad d_1, d_0} & F(X_0)\lto<0.75ex>^{\qquad s_0} .}\enddiagram $$ Note that for $n>1,$ higher levels of ${\bf{F}}^{(1)}$ are generated by the degenerate elements. \begin{lem}\label{ber} We assume given the 1-skeleton ${\bf{F}}^{(1)}$. Let $d_0^1$ and $d_{1\text{ }}^1$be evaluation homomorphisms. Then i) \ \ {\rm Ker}$d_0^1= {\langle Y_1\rangle},$ ii) \ {\rm Ker}$d_1^1={\langle Z\rangle },$\\ where $Z=\{s_1(y)^{-1}s_0(y) : y \in Y_1\}$ and ${\langle Y_1\rangle}$ is normal closure of $Y_1.$ \end{lem} \begin{pf} Clear. \end{pf}\\ Note $\pi_0({\bf{F} }^{(1)})\cong G $. The link between the bottom step of a step-by-step construction of the resolution and that of a CW-basis $\mathfrak{F}$ is thus clear. The 2-skeleton gives the non-degenerate elements of the resulting CW-basis, $\mathfrak{F}_2,$ and in general we can take $Y_n\cong \Omega^{n-1},$ and $\mathfrak{F}_n=Y_n\cup~\bigcup s_i(\mathfrak{F}_{n-1}).$ For both combinatorial and computational purposes, the way in which $Y_n$ corresponds to $\Omega^{n-1}$ can be important and in general it is necessary to specify the function $ g^{n-1}:\Omega^{n-1}\longrightarrow NF_n $ or its last face $ d_ng^{n-1}:\Omega^{n-1}\longrightarrow NF_{n-1}. $ {\bf Remark:} For homological and computational reasons, it is often useful also to specify the contracting homotopyon the underlying simplicial set of ${\mathbf F}$ and to build this into the resolution progresses. We will not discuss how to do this here however as it is not needed for our immediate purposes. \medskip Before carrying on the `step-by-step' construction of the free simplicial group, we will interpret the first homotopy group, $\pi _1({\bf {\mathbf F}}^{(1)})$, of ${\mathbf F}^{(1)}$ to find what it looks like. For any simplicial group ${\mathbf F}$, if ${\mathbf F}={\mathbf F}^{(1)},$ then, $$ \pi_1({\mathbf F})={\rm Ker}(d_1 :{\rm Ker}d_0^1/[{\rm Ker}d_1^1, ~{\rm Ker}d_0^1] {\longrightarrow F_0}). $$ Indeed, by definition, the first homotopy group is $$ \pi_1({\mathbf F})=(\text{Ker}d_0^1\cap \text{Ker}d_1^1)/d_2^2(\text{Ker} d_0^2\cap \text{Ker}d_1^2). $$ By a lemma of Brown and Loday \cite{bl}, see also \cite{mp1}, the denominator of this homotopy group isexactly $$ \partial_2(NF_2)=d_2^2(\text{Ker}d_0^2\cap \text{Ker}d_1^2)=[\text{Ker}d_0^1, ~\text{Ker}d_1^1] $$ and the morphism $$ \delta :\text{Ker}d_0^1/\partial _2(NF_2)\longrightarrow F_0, $$ where $\delta =d_1$ (restricted to $NF_1/\partial _2NF_2)$, is a crossed module. Here $NF_0$ acts on $NF_1/\partial _2NF_2$ by conjugation via $% s_0,$ that is, $$ \begin{array}{ccl} NF_1/\partial_2NF_2\times NF_0 & \longrightarrow & NF_1/\partial _2NF_2, \\ ({x},\overline{y}) & \longmapsto & {}^{x} \overline{y}=\overline{s_0(x)ys_0(x)^{-1}}, \end{array} $$ where $\overline{y}$ denotes the corresponding element of $NF_1/\partial_2NF_2$ whilst $y\in NF_1.$ \begin{eqnarray*} \pi _1({\mathbf F}) & = &\text{Ker}(\text{Ker}d_0^1/\partial_2(NF_2)\longrightarrow F_0)\\ & = &\text{Ker}(NF_1/[\text{Ker}d_1, ~\text{Ker}d_0]\longrightarrow F_0). \end{eqnarray*} \begin{prop} Given a presentation $P=\langle{X_0~|~R\rangle}$ of a group $G$ and $ {\mathbf F}^{(1)}$, the 1-skeleton of the free simplicial group generated by this presentation, then $$ \delta :NF_1^{(1)}/\partial_2(NF_2^{(1)})\longrightarrow NF_0^{(1)} $$ is the free crossed module on $R\rightarrow {F(X_0)},$ and $\pi_2({\mathbf F}^{(1)})$ is the module of identities of the presentation $P.$ \end{prop} \begin{pf} Clear. \end{pf} \medskip Note that for the case of ${\mathbf F}^{(2)}$, if $x_i,~x_j$ are in $NF_1^{(2)},$ then generators of the normal subgroup $NF_2^{(2)}\cap D_2$ are of the form $ [s_1(x_i)^{-1}s_0(x_j),~s_1(x_j)]. $ We now will recall the next step of the construction of a free simplicial group. We take a set of generators $ \Omega ^1=\{S_i\}\subset \pi _1({\mathbf F}^{(1)}) $ and kill off the elements in the homotopy group $\pi _1({\mathbf F}^{(1)})$ by adding new indeterminates $Y_2=\{y_i\}$ into $F_2^{(1)}$ where $Y_2$ is in $1-1$ correspondence with $\Omega ^1$ to establish $$ F_2^{(2)}=F_2^{(1)}(Y_2)=F(s_1s_0(X_0)\cup s_0(Y_1)\cup s_1(Y_1)\cup Y_2) $$ together with $$ d_0^2(y_i~)=1,\quad d_1^2(y_i~)=1,\quad d_2^2(y_i)=S_i,\textrm{ mod }\partial_3NF_3^{(2)}. $$ Hence the 2-skeleton ${\mathbf F}^{(2)}$ looks like $$ \diagram{ F(s_1s_0(X_0)\cup s_0(Y_1)\cup s_1(Y_1)\cup Y_2)\ar[r]<0.25ex> \ar[r]<1ex> \ar[r]<1.75ex>^{\qquad\qquad d_0 ,d_1 ,d_2 }&F(s_0(X_0)\cup (Y_1)) \ar[l]<0.75ex> \ar[l]<1.50ex>^{\qquad\qquad s_1,s_0}\ar[r]<0.25ex>\ar[r]<1ex>^{\qquad d_1, d_0} & F(X_0)\ar[l]<0.75ex>^{\qquad s_0} ,}\enddiagram $$ and, of course, $$ \begin{array}{lll} F_2&=&F(s_1s_0(X_0)\cup s_0(Y_1)\cup s_1(Y_1)\cup Y_2)\\ &\cong&F(s_1s_0(X_0))\ast F(s_0(Y_1))\ast F(s_1(Y_1))\ast F(Y_2). \end{array} $$ In ${\bf F}^{(2)}$, higher levels than dimension 2 are generated by degenerate elements. This pattern, of course, continues to higher dimensions. We thus have in each dimension, $k$-dimensional construction data $(\Omega^{k}, g^{(k)})$ and a $k^{th}$-level presentation of the group, $G.$ The various $\Omega^{k}$ thus provide us with a $CW$-basis for ${\bf F}.$ \section{Free Crossed Resolutions} In this section we want to examine in slightly more detail this step-by-step construction through the perspective of the corresponding crossed complex, examining not only to see if $\mathfrak{C}(\mathbb{F})$ is a crossed resolution of a group $G,$ but also how the homotopy type of $\mathfrak{C}(\mathbb{F}^{(k)})$ is constructed from $\mathfrak{C}(\mathbb{F}^{(k-1)}).$ Knowledge of this process would seem essential if the construction of crossed resolutions is to be `mechanised'. It also helps in the interpretation of homological invariants and their linkage with combinatorial properties of a presentation or of a higher level presentation of a group $G.$ As the analysis is applicable in greater generality, we start by looking at an arbitrary free simplicial group with chosen $CW$-basis. A `step-by-step' construction of a free simplicial group is constructed from simplicial group inclusions $$ {\bf F}^{(0)}\subseteq \text{ }{\bf F}^{(1)}\text{ }\subseteq \text{ }{\bf F}^{(2)}\text{ }\subseteq \text{ }\ldots $$ We take the functor $\mathfrak{C}$ which is described in Section~\ref{fa}, to see what ${C}_n({\bf F}^{(k)})$ looks like, where ${\bf F}^{(k)}$ is the k-skeleton of that construction, concentrating our attention in low dimensions. For $k=0,$ we have the 0-skeleton ${\mathbb F}^{(0)}$ of the construction% $$\diagram {\mathbb F}^{(0)}:\ldots \quad {F(X_0)}\rto<.3ex>\rto<-.3ex>& {F(X_0)}\rightarrow {F(X_0)}/N. \enddiagram $$ Here ${\bf F}^{(0)}$ is the trivial simplicial group in which in every degree $n,$ $F_n^{(0)}= {F(X_0)}$ and $d_i^n=\ $id$\ =s_j^n.$ It is easy to see that ${C}_0({\bf{F}}^{(0)})= {F(X_0)}$ as $NF_1\cap D_1$ is trivial. \noindent The 1-skeleton is $$ \diagram{ ...{~}{F}(s_1s_0(X_0)\cup s_0(Y_1)\cup s_1(Y_1)) \rto<0.25ex> \rto<1ex> \rto<1.75ex>^{\qquad\qquad d_0 ,d_1 ,d_2 } & {F}(s_0(X_0)\cup Y_1) \lto<0.75ex> \lto<1.50ex>^{\qquad\qquad s_1,s_0} \rto<0.25ex> \rto<1ex>^{\qquad d_1, d_0} & {{F}(X_0)} \lto<0.75ex>^{\qquad s_0},}\enddiagram $$ and since $F_2^{(1)}$ is generated by degenerate elements, $% F_2^{(1)}=D_2$, so the crossed complex term ${C}_1({\bf F}^{(1)})$ is the following% $$ \begin{array}{llll} C_1({\bf F}^{(1)}) & = & \dfrac{NF_1^{(1)}}{(NF_1^{(1)}\cap D_1) \partial _2(NF_2^{(1)}\cap D_2)}, & \\ \\ & = & \dfrac{NF_1^{(1)}}{\partial _2(NF_2^{(1)}\cap D_2)}\quad & \text{since }NF_1\cap D_1=1, \\ \\ & = & \dfrac{NF_1^{(1)}}{\partial _2(NF_2^{(1)})}\quad & \text{as }F_2^{(1)}=D_2. \end{array} $$ By Lemma~\ref{ber} and the Brown-Loday lemma \cite{bl}, we have $NF_1^{(1)}= \langle Y_1\rangle$ and $\partial _2(NF_2^{(1)})$ is generated by the Peiffer elements, respectively. It then follows that $$ C_1({\bf F}^{(1)})=\langle Y_1\rangle /P_1.\ $$ Here $P_1$ is the first dimensional Peiffer normal subgroup. The proof of Theorem~\ref{rin} from \cite{bh3} interprets within this context as showing that $ \partial _1:\langle Y_1\rangle/P_1\longrightarrow {F(X_0)} $ is the free crossed module on the presentation $\langle X_0\mid d_1(Y_1)\rangle$ of $\pi_0({\bf F}).$\\ Looking at the case $2,$ the 2-skeleton of the construction is% $$ \diagram{ ...{~}{F}(s_1s_0(X_0)\cup s_0(Y_1)\cup s_1(Y_1)\cup Y_2) \rto<0.25ex> \rto<1ex> \rto<1.75ex>^{\qquad\qquad d_0 ,d_1 ,d_2 } & {F}(s_0(X_0)\cup Y_1) \lto<0.75ex> \lto<1.50ex>^{\qquad\qquad{~}~ s_1,s_0} \rto<0.25ex> \rto<1ex>^{\qquad d_1, d_0} & {{F}(X_0)} \lto<0.75ex>^{\qquad s_0}.}\enddiagram $$ As before $F_3^{(2)}=D_3$ as $F_3^{(2)}$ is generated by the degeneracy elements. Thus the second term of the crossed complex is $$ \begin{array}{lllc} C_2({\bf F}^{(2)}) & = & \dfrac{NF_2^{(2)}}{(NF_2^{(2)}\cap D_2)\partial _3(NF_3^{(2)}\cap D_3)}, & \\ \\ & = & \dfrac{NF_2^{(2)}}{(NF_2^{(2)}\cap D_2)\partial _3(NF_3^{(2)})}\quad & \text{as }F_3^{(2)}=D_3. \end{array} $$ If $x,y\in NF_1,$ then $NF_2\cap D_2$ is generated by the elements of the form $ [s_1x^{-1}s_0x,~ s_1y] $ and in general, if $x,~y\in NF_{n-1},$ then $ s_{n-1}x^{-1}s_{n-2}x,~ s_{n-1}y] \in NF_n\cap D_n. $ Now look at $\partial _3(NF_3^{(2)})$ in terms of the skeleton ${\bf F} ^{(2)}.$ In a similar way to the proof of \thinspace Lemma~\ref{ber} and as $d_0^2(y_i)=d_1^2(y_i)=1 \ y\in Y_2$, \ one can readily obtain that: $$ NF_2^{(2)}= \langle s_1(Y_1)\cup Y_2\rangle\cap\langle Z\cup Y_2\rangle. $$ where $Z$ as in Lemma~\ref{ber}. On the other hand, \cite{mp1} shows that on writing $K_I = \bigcap_{i\in I} \text{Ker}d_i$ for $I \subseteq [n-1]$, $$ \partial _3(NG_3^{(2)})=\prod_{I,J} [K_I, ~K_J] [K_{\{0,2\}}, ~K_{\{0,1\}}] [K_{\{1,2\}}, ~K_{\{0,1\}}] [K_{\{1,2\}}, ~K_{\{0,2\}}], $$ where $I\cup J=[2],~I\cap J=\emptyset $, so this is generated by the following elements: for $x_i\in NF_1=\text{Ker}d_0$ and $y_1,y_2\in NF_2=\,$Ker$d_0\cap $Ker$d_1$ with $1\leq i,j\leq n$, $$ \begin{array}{rcl} \lbrack s_0{x_i}^{-1}s_1s_0d_1x_i,~ y_1 \rbrack &\quad(1)\\ \lbrack s_1{x_i}^{-1}s_0x_i,~ s_1d_2(y_1){y_1}^{-1} \rbrack &\quad(2) \\ \lbrack x_is_1d_2{x_i}^{-1}s_0d_2x_i,~ s_1y_1 \rbrack &\quad(3)\\ \lbrack {y_1}^{-1}s_1d_2y_1,~ y_2 \rbrack &\quad(4)\\ \lbrack y_1s_1d_2{y_1}^{-1}s_0d_2y_1,~ y_2 \rbrack &\quad(5) \\ \lbrack y_1s_1d_2{y_1}^{-1}s_0d_2y_1,~ s_1d_2(y_2){y_2}^{-1} \rbrack &\quad(6). \end{array} $$ The normal subgroup generated by these elements will be denoted by $P_2$ and will be called the second dimensional Peiffer normal subgroup. We thus in principle have not only an explicit presentation of $C_2({\bf F}^{(2)})$ but a list of seven `generic' moves analogous to the Peiffer moves introduced by Brown and Huebschmann, \cite{bh3}. Writing $Q_2=NF_2^{(2)}\cap D_2,$ we get the second term of the crossed complex as follows $$ C_2({\bf F}^{(2)})=\frac{\langle s_1(Y_1)\cup Y_2\rangle \cap\langle Z\cup Y_2\rangle}{Q_2 \cdot P_2} $$ \begin{prop} Let ${\bf F}^{(2)}$ be the 2-skeleton of a free simplicial group resolving $G= F(X_0)/N.$ Then $$ {\mathfrak{C}}^{(2)}:\quad NF_2^{(2)} /(Q_2 \cdot P_2)\stackrel{\partial _2% }{\rightarrow }\langle Y_1\rangle /P_1 \stackrel{\partial _1}{\rightarrow }{F(X_0)}\stackrel{g}{ \rightarrow }{F(X_0)}/{N}\stackrel{f}{\rightarrow }1 $$ is the 2-skeleton of a free crossed resolution of $G$ where $\partial _2$ and $\partial _1$ are given respectively by: for $y_1\in\langle s_1(Y_1)\cup Y_2\rangle \cap \langle Z\cup Y_2\rangle$ and $x_i\in \langle Y_1\rangle,$ $$ \partial _2(y_1(Q_2\cdot P_2))=d_2(y_1) P_1\text{ and }\partial _1(x_i P_1)=d_1(X_i). $$ where $NF_2^{(2)}$ is $(\langle s_1(Y_1)\cup Y_2\rangle\cap\langle Z\cup Y_2\rangle)$. \end{prop} \begin{pf} This follows immediately from the description of the `step-by-step' construction of the free simplicial group. \end{pf} \medskip This result gives a combinatorial description of the $C_2({\bf F}^{(2)})$ term and if we manipulate the elements of $s_1(Y_1)\cup Y_2$ and $Z\cup Y_2$, remembering that $Z=\{s_1(y)^{-1}s_0(y): y\in Y_1\},$ we can identify the generators as elements in the module of identities of the presentation $\langle X_0\mid d_1(Y_1)\rangle.$ The elements of $Y_2$ map via $d_2$ to a set of generators of this module since, of course, that is how they were chosen. To complete our analysis of the r\^ole of a CW-basis in a free simplicial resolution ${\mathbb F} = ({\bf F}, g)$ of a group $G$, we need to check that $(C({\bf F}),C(g))$ is a free {\em crossed} resolution of $G$ and to see what happens to the CW-basis in the `conversion'. First a proposition showing how homotopies behave under the functor, $\mathfrak C$. \begin{prop} Suppose ${\mathbf f_0}, {\mathbf f_1} :{\mathbf G}\rightarrow {\mathbf H}$ are morphisms of simplicial groups and ${\mathbf h} : {\mathbf f_0}\simeq{\mathbf f_1}$ is a homotopy between them. Then ${\mathbf h}$ induces a homotopy ${\mathfrak C}({\mathbf h}) : \pi(1) \otimes{\mathfrak C}({\mathbf G}) \rightarrow{\mathfrak C}({\mathbf H})$ between ${\mathfrak C}{(\mathbf f_0})$ and ${\mathfrak C}({\mathbf f_1})$. \end{prop} (Here the $\otimes $ is the tensor product of crossed complexes introduced by Brown and Higgins, \cite{bh2}, and $\pi(1) := \pi(\Delta^1)$ is the groupoid `unit interval'. For more on the homotopy theory of crossed complexes, the simplicial category theory of the category of crossed complexes, etc., see \cite{bgpt} and \cite{tonks}.) \begin{pf} The homotopy ${\mathbf h}$ can be realised as a morphism, ${\mathbf h} : \Delta[1] \bar{\otimes} {\mathbf G} \rightarrow {\mathbf H},$ where $\Delta[1] \bar{\otimes} \quad $ is the simplicial tensor within the simplicially enriched category of simplicial groups (or groupoids) (see Quillen, \cite{q}, or the discussion in \cite{mp}.) This is given as a colimit of copies of ${\bf G}$ by the construction outlined in \cite{q}. The functor ${\mathfrak C}$ can be thought of in two equivalent ways. It is either the composite of the reflection onto the variety of simplicial group(oid) T-complexes (cf. \cite{ep}) followed by the equivalence between that and the category of crossed complexes, or alternatively it uses the Cegarra-Carrasco equivalence between simplicial groupoids and hypercrossed complexes of group(oid)s followed by the reflection onto the variety of crossed complexes within that category. (The advantage at this point in using groupoids is that $\pi(1)$ is naturally a groupoid, but this can be avoided if desired.) From either description it is clear that ${\mathfrak C}$ will preserve colimits and thus tensors with simplicial sets, thus $${\mathfrak C}(\Delta[1]\bar{\otimes} {\mathbf G}) \cong \Delta[1]\bar{\otimes}{\mathfrak C}({\mathbf G})\cong \pi(1)\otimes {\mathfrak C}({\mathbf G}).$$ Composing ${\mathfrak C}({\mathbf h})$ with these isomorphisms gives the result.\end{pf} \begin{cor} If ${\mathbf g} : {\mathbf F} \rightarrow {\mathbf K}(G,0)$ is a free simplicial resolution of $G$, then ${\mathfrak C}({\mathbf g}) : {\mathfrak C}({\mathbf F}) \rightarrow {\mathfrak C}({\mathbf K}(G,0))= G$ is a free crossed resolution of $G$. \end{cor} \begin{pf} The data on ${\mathbf g}$ can be specified by giving a homotopy between the identity on ${\mathbf F}$ and the map that `squashes $NF$ down to $G$' and then uses a section of the augmentation map, $g_0$, to yield a map back to $NF_0$. The corollary now follows from the previous result applied to this simplicial homotopy. \hfill\end{pf} \medskip To finish the comparison, we will show that each $C_n({\mathbf F})$ is a free $G$-module on $Y_n$ if $n\geq 2$. We start with $n = 2$ but in fact almost the same proof works in higher dimensions. Suppose that $M$ is a $G$-module and $\Theta :Y_2 \rightarrow M$ is a function, we want to prove that $C_2({\mathbf F})$ is free on $Y_2$, so we need to extend $\Theta$ to a map on $C_2({\mathbf F})$. Form the crossed complex $$\ldots \rightarrow M \rightarrow 1\rightarrow G$$ with $M$ in dimension 2, $G$ in dimension 0, all other levels being trivial and the action of $G$ on $M$ being the given one. This has an associated simplicial group $S(M,G)$ with $NS(M,G)$ this crossed complex. There is an obvious morphism, $\phi$ from ${\mathbf F}^{(1)}$ to $S(M,G)$, inducing the quotient morphism $g : F(X_0) \rightarrow F(X_0) / N \cong G$. As $\pi_1(S(M,G))$ is trivial, Proposition \ref{free} applies to show $\phi$ extends over ${\mathbf F}^{(2)}$ also extending $\Theta$. Now we use ${\mathfrak C}$ to pass back to crossed complexes to get $${\mathfrak C}(\phi) : {\mathfrak C}({\mathbf F}^{(2)}) \rightarrow M$$ extending $\Theta$. As $C_2({\mathbf F}^{(2)}) \simeq C_2({\mathbf F}) $, this proves the claim that $C_2({\mathbf F}^{(2)}) $ is a free $G$-module on $Y_2$. Of course, the only difference that is needed in dimension $n$ is in the definition of $S(M,G)$, where $M$ is placed in dimension $n$ and $\Theta : Y_n \rightarrow M$ is given. We have proved: \begin{prop} If $\mathbb{F}$ is a simplicial resolution of $G$ given by a construction data sequence $\{(Y_i, g^{(i)}), i=0,1,\ldots\}$ and ${\mathbf F}^{(k)}$ is the corresponding $k$-skeleton, then if $k\geq 2,$ $C_k(\mathbb{F}^{(k)})$ is a free $G$-module on $Y_k.$ \end{prop} Summarising we get: \begin{thm} The `step-by-step' construction of simplicial resolution of a group, $G$, yields a `step-by-step' construction of a crossed resolution of $G$ via the crossed complex construction, $\mathfrak{C}.$ \end{thm} As a bonus for our method we also have given an explicit description of the crossed complex construction in low dimensions. The construction data to dimension $n$ yields an $n$-dimensional word system in the sense of R.A.Brown. What is less clear, as we have mentioned before, is why the word system given by Whitehead (see \cite{rab}, Example 2.2.3) would not seem to lift back to give construction data for a free simplicial group.
\section{Introduction} It is obviously the most challenging problem how to formulate the quantum gravity and the standard model in a unifying and constructive way. Towards a possible solution to this problem, the current trend is heading to the string related topics\cite{string1}\cite{string2}. It is, however, not obvious that the string is the only possible formulation leading to the unified theory including quantum gravity. In fact, the two dimensional quantum gravity was formulated by a lattice gravity, the dynamical triangulation of random surface\cite{DT} which was analytically confirmed by Liouville theory\cite{KPZ}. On the other hand the three dimensional Einstein gravity was successfully formulated by the Chern-Simons action even at the quantum level \cite{Witten1}. There are thus other formulations of quantum gravity than the string related formulation. One of the important motivations that the super string could be the genuine formulation of unifying all the forces of gauge theories is that the super string may be able to control the divergences even with gravity, and thus the renormalizability and unitarity are natural consequences of the formulation. An alternative formulation to control the divergences would be the lattice formulation. Suppose we aim to formulate a unified model, what could be the possible criteria to believe that it could be a realistic model. Eventually the unified model should explain the origin of the following phenomenological parameters and characteristics: \begin{enumerate} \item The group structure of the standard model: $SU(3)\times SU(2)\times U(1)$. \item The number of generation = 3. \item The pairing structure of quarks and leptons in the standard model: Quarks interact strongly, weakly and electromagnetically while leptons interact differently. \item Our space-time is four dimension and Minkowskian. \item The quark and lepton masses, the mixing parameters; Cabbibo-Kobayashi-Maskawa angles and possible lepton mixing angles. \end{enumerate} The fundamental unified theory should eventually propose a mean to evaluate the item 5 quantitatively. From the experiences of the lattice QCD it would be difficult to calculate the phenomenological quantities analytically from the fundamental theory. Instead we need to evaluate them numerically thus we need to formulate a constructive definition of a regularized unified theory including gravity. Concerning to the issues of quantum gravity it is more difficult to judge what could be the experimental evidences to confirm the quantum nature of the gravity. It could possibly be reflected to the large scale structure of the universe yet unconvincing. As we show later the quantum gravity in two dimensions can be well understood by the numerical simulations on the lattice which are confirmed analytically as well. We expect that numerical method by lattice would be the only mean to evaluate the quantum nature of gravity even in higher dimensions. We thus believe that lattice theory is again a good candidate to fulfill our requirements for the quantitative unified theory. In the above phenomenologically known results the first four items could be understood easier than the last one and could be related with the super symmetry. In this summary of overview we propose the generalized gauge theory which was proposed previously by the present author and Y.Watabiki \cite{KW2} as a formulation towards a unified model on the lattice. In order to persuade the readers to accept the ideas and formulation I will collect the suggestive known results and include several recent investigations towards this direction and thus the summary is aimed to be self-contained. Our formulation is a non-standard approach towards the unified model and the formulation is not yet completed but there are several hopeful evidences that this approach may play an important role for the unified theory of the fundamental interactions. Here we list the contents of this summary towards a non-standard approach of a unified model on the lattice. \begin{flushleft} 1. Introduction \\ 2. Suggestive Known Results towards Unified Model on the Lattice \\ ~~ 2.1 Fermionic Matter on the Lattice \\ ~~ 2.2 Success of Two Dimensional Quantum Gravity on the Lattice \\ ~~~~ 2.2.1 Microscopic Description of Two Dimensional Random Surface \\ ~~~~ 2.2.2 Numerical Results on the Fractal Structure of Two Dimensional Quantum \\ \hspace{1.6 cm} Gravity on the Lattice \\ ~~ 2.3 Susskind Fermion---Staggered Fermion---Dirac K$\ddot{\hbox{a}}$hler Fermion on the Lattice \\ ~~ 2.4 Chern-Simons Gravity and Ponzano-Regge Model \\ ~~~~ 2.4.1 Chern-Simons Gravity \\ ~~~~ 2.4.2 Ponzano-Regge Model \\ ~~ 2.5 Four Dimensional Gravity on the Lattice \\ 3. A Possible Formulation towards Gauge Gravity coupled Matter on the \\ \hspace{0.5 cm} Simplicial Lattice Manifold \\ ~~ 3.1 Possible Formulations towards Unified Model on the Lattice \\ ~~ 3.2 Generalized Gauge Theory \\ 4. Gravity on the Lattice \\ ~~ 4.1 First Step towards the Generalized Chern-Simons Actions on the Lattice \\ ~~ 4.2 Lattice Chern-Simons Gravity \\ ~~~~ 4.2.1 Gauge Invariance on the Lattice \\ ~~~~ 4.2.2 Calculation of Partition Function \\ ~~~~~~ 4.2.2.1 $e$ integration \\ ~~~~~~ 4.2.2.2 $U$ integration \\ ~~~~ 4.2.3 The Continuum Limit of the Lattice Chern-Simons Gravity \\ 5. Quantization of Generalized Gauge Theory \\ 6. Generalized Yang-Mills Theory \\ ~~ 6.1 Generalized Topological Yang-Mills Theory \\ ~~~~ 6.1.1 Instanton Gauge Fixing of Topological Yang-Mills Model \\ ~~~~ 6.1.2 Twisted $N=2$ Super Yang-Mills Action with Dirac-K$\ddot{\hbox{a}}$hler Fermions \\ ~~ 6.2 Weinberg-Salam Model from Generalized Gauge Theory \\ ~~~~ 6.2.1 Generalized Gauge Theory with Dirac-K$\ddot{\hbox{a}}$hler Fermions \\ ~~~~ 6.2.2 Weinberg-Salam Model from Generalized Gauge Theory as \\ \hspace{1.6 cm} Non-Commutative Geometry Formulation \\ 7. Possible Scenario and Conjectures for the Unified Model on the Lattice \\ \end{flushleft} \setcounter{equation}{0} \section{Suggestive Known Results towards Unified Model on the Lattice} \subsection{Fermionic Matter on the Lattice} In considering the formulation towards the unified model on the lattice, I will first give a very suggestive and simple example to figure out how to find the possible model on the lattice. The first example is the two dimensional Ising model. It is well known that the Ising model has a second order phase transition point and the model leads to a free fermion theory in a flat space at the phase transition point in the continuum limit. \begin{figure} \begin{center} \begin{minipage}[b]{0.3\textwidth} \epsfxsize=\textwidth \epsfbox{ising.eps} \end{minipage} \hspace{1 cm} \begin{minipage}[b]{0.3\textwidth} \epsfxsize=\textwidth \epsfbox{isinggr.eps} \end{minipage} \end{center} \caption{Ising spin on the square lattice and dynamically triangulated lattice} \label{fig:ising} \end{figure} In this system the lattice is the two dimensional square lattice and the matter field is sitting at the sites of the lattice and takes value $\pm 1$. What is surprising here is that the simple square lattice with the simplest matter at the site reproduces the fermionic matter in the continuum limit. It is by now established that the Ising spin on the dynamically triangulated lattice reproduces the free fermion coupled to a gravitational background in the continuum limit. See fig.\ref{fig:ising}. This is a very symbolic example that matter fermion is generated via degrees of freedom of field sitting on sites and the curved space-time background is generated by the dynamically triangulated lattice in the continuum limit. In other words field theoretical matter and background gravitational field are essentially reproduced by the lattice formulation. It is important to recognize that there is a microscopic formulation to see how the fermionic degrees appear at the lattice level\cite{Ising1}\cite{Dotsenko}. Let me sketch the formulation. The partition function of Ising model is \begin{equation} Z=\sum_{\sigma=\pm} \exp\{ \beta\sum_{<ij>}\sigma_i\sigma_j\}, \end{equation} where $<ij>$ is a nearest neighbor pair of sites on a two dimensional square lattice. We introduce the so called disorder parameter\cite{Kadanoff} \begin{equation} \mu_x = \prod^x_{-\infty} \exp {\{-2\beta \sigma\sigma'\}}, \end{equation} where the exponents in the product correspond to the links which are crossed by the dashed line starting from the dual site $x$ and ending at $-\infty$. We introduce the product of the disorder variable $\mu_x$ and order variables $\sigma_x$ \begin{eqnarray} \begin{minipage}[c]{4.0cm} \epsfxsize=\textwidth \epsfbox{dual1.eps} \end{minipage} ~~~~~~=~~&\mu_x\sigma_{x_1}~~\equiv~~ \chi^1(x) \\ \begin{minipage}[c]{4.0cm} \epsfxsize=\textwidth \epsfbox{dual2.eps} \end{minipage} ~~~~~~ =~~&\mu_x\sigma_{x_2}~~\equiv~~ \chi^2(x) \\ \begin{minipage}[c]{4.0cm} \epsfxsize=\textwidth \epsfbox{dual3.eps} \end{minipage} ~~~~~~=~~&\mu_x\sigma_{x_3}~~\equiv~~ \chi^3(x) \\ \begin{minipage}[c]{4.0cm} \epsfxsize=\textwidth \epsfbox{dual4.eps} \end{minipage} ~~~~~~=~~&\mu_x\sigma_{x_4}~~\equiv~~ \chi^4(x), \end{eqnarray} where $x_1$, $x_2$, $x_3$ and $x_4$ are original sites and surround the dual site $x$. Here we assume that the $\sigma\mu$ variables are always inside some correlation function $\chi^\alpha(x)=\mu_x\sigma_{x_\alpha}\sim<\mu_x\sigma_{x_\alpha}\cdots>$. Then the effect of the disorder variable is to flip the sign of $\beta\sigma\sigma'$ in the action along the dashed line. There is the following identity: \begin{equation} \sigma\exp \{-2\beta\sigma\sigma'\}=ch(2\beta)\sigma -sh(2\beta)\sigma', \end{equation} which combines with the disorder variable $\mu_x$ and leads to a graphical relation \begin{equation} \begin{minipage}[c]{2.0cm} \epsfxsize=\textwidth \epsfbox{dual5.eps} \end{minipage} ~~=~~ch(2\beta)~ \begin{minipage}[c]{2.0cm} \epsfxsize=\textwidth \epsfbox{dual6.eps} \end{minipage} ~-~sh(2\beta) \begin{minipage}[c]{2.0cm} \epsfxsize=\textwidth \epsfbox{dual7.eps} \end{minipage} , \end{equation} equivalently \begin{equation} \chi^1(x)=ch(2\beta)\chi^2(x-\hat{1}) -sh(2\beta)\chi^3(x-\hat{1}), \end{equation} where $\hat{1}$ is a unit vector of 1 direction on the lattice. We can obtain similar relations for $\chi^2(x)$, $\chi^3(x)$ and $\chi^4(x)$, where we can use the invariance under a simultaneous change of $\sigma_n \rightarrow -\sigma_n$ and $\beta\sigma\sigma_n \rightarrow -\beta\sigma\sigma_n$ for an arbitrary given site $n$. Redefining $\chi^\alpha(x)$, \begin{eqnarray} \psi^1(x)&=\frac{1}{2}(\chi^2(x)+\chi^3(x)), ~~~ \psi^2(x)&=\frac{1}{2}(\chi^1(x)+\chi^2(x)), \\ \psi^3(x)&=\frac{1}{2}(-\chi^4(x)+\chi^1(x)), ~~~ \psi^4(x)&=\frac{1}{2}(-\chi^3(x)-\chi^4(x)), \end{eqnarray} we obtain \begin{equation} \left[\begin{array}{c} \psi^1(x)\\ \psi^2(x)\\ \psi^3(x)\\ \psi^4(x) \end{array} \right] =e^{-2\beta} \left[\begin{array}{cccc} 1&1&0&-1\\ 1&1&1&0\\ 0&1&1&1\\ -1&0&1&1 \end{array} \right] \left[\begin{array}{c}\psi^1(x-\hat{1})\\ \psi^2(x-\hat{2})\\ \psi^3(x-\hat{3})\\ \psi^4(x-\hat{4}) \end{array} \right], \end{equation} where the unit vectors are related $-\hat{3}=\hat{1}$ and $-\hat{4}=\hat{2}$. This can be understood as a hopping relation of the $\psi^\alpha$ fields on the lattice. Under the proper renormalization of the $\psi^\alpha$ fields, the hopping matrix is identified with the inverse propagator of fermion fields\cite{Ising1}. It is important to recognize that the microscopic description of the fields on the lattice is reflected to the field theoretic description of fermionic field in the continuum limit. By now we know that the Ising spin located at sites of a dynamically triangulated lattice leads to a fermionic matter coupled to gravity in the continuum limit\cite{Migdal}\cite{Kazakov}. In this case there is also a microscopic description of how fermionic degrees appear on the random lattice. Let me briefly sketch the derivation. Here we don't follow to the matrix model formulation\cite{Kazakov} but give more direct derivation. We first note the standard observation concerning Ising model: the general configuration of spins can be considered as a set of alternating black drops ($\sigma=1$) and white bubbles ($\sigma=-1$). The boundaries between drops and bubbles are drawn on the dual graph. See fig. \ref{fig:isinggr1} . \begin{figure} \begin{center} \begin{minipage}[b]{0.3\textwidth} \epsfxsize=\textwidth \epsfbox{isinggr2.eps} \end{minipage} \end{center} \caption{Ising spin on a dynamically triangulated lattice} \label{fig:isinggr1} \end{figure} The dual graph includes only three vertex on a dual site and thus coincides with Feynmann graphs of $\phi^3$ theory. If we look at particular dual site, a boundary line may pass through the site or may not touch at all. We can show that the above mentioned property is reproduced by the following fermionic action on the lattice\cite{Migdal} \begin{equation} Z=\int \prod_i d^2 \psi(i) exp(-S), \end{equation} with the action \begin{equation} S=\frac{1}{2}\sum_i \bar{\psi}(i)\psi(i)- \frac{1}{2}\sum_{<ij>}\bar{\psi}(i)K(i,j)\psi(j). \end{equation} We adopt the notation $\bar{\psi}_\beta=\psi_\alpha\epsilon_{\alpha\beta}$ and the normalization of Grassmann integral \begin{equation} \int d^2\psi \psi_\alpha\bar{\psi}_\beta = \delta_{\alpha\beta}. \end{equation} There are two important properties to be satisfied by the action. (1) The boundary line does not possess particular direction thus the fermion must have Majorana nature. (2) The back-tracking motion is absent. These properties are satisfied by the following constraints: \begin{eqnarray} (1)&~~~~\bar{\psi}(i)K(i,j)\psi(j)&= \bar{\psi}(j)K(j,i)\psi(i) \\ (2)&~~~~K(i,j) K(j,i) &= 0. \end{eqnarray} The Majorana condition (1) can be written in a matrix form: \begin{equation} \sigma_2 K(i,j)= -(\sigma_2 K(j,i))^T, \end{equation} where $(\sigma_2)_{\alpha\beta}=i\epsilon_{\alpha\beta}$ is the Pauli matrix and $T$ stands for transpose. The above two constraints will be satisfied, provided $K(i,j)$ has the form \begin{equation} K(i,j)=A(ij)({\bf 1} + \sigma_1 n_1(ij) + \sigma_3 n_3(ij)), \end{equation} where \begin{equation} A(ij)=A(ji),~~~n_1(ij)=-n_1(ji),~~~n_3(ij)=-n_3(ji),~~~ (n_1(ij))^2+(n_3(ij))^2=1. \end{equation} We can then guess a possible strategy how to get a field theory coupled gravity on the lattice. The gravitational background is generated by the dynamical triangulation and the matter fields are generated by the fields on the simplex of the simplicial manifold. \subsection{Success of Two Dimensional Quantum Gravity on the Lattice} As we have seen in the previous section that Ising model on the two dimensional square lattice is equivalent to the free fermion theory of the flat space time in the continuum limit. On the other hand it is analytically known by now that Ising model on the randomly triangulated lattice is equivalent to the free fermion theory coupled to the two dimensional gravity\cite{Migdal}\cite{Kazakov}. These examples suggest that curved space time is generated by the random lattice and the matter fields are induced by some degrees of freedom on a simplex: for Ising model $\pm 1$ values on the sites (0-simplex). In two dimensions the relation between the lattice theory and the corresponding continuum field theory were analytically established with the help of conformal field theory. The central charge $c$ specifying a matter content of the continuum theory is a measure to differentiate different types of matrix models. The equivalence of a lattice model and the corresponding matrix model is established and the Liouville theory gave complementary understandings to the lattice theories\cite{KPZ}\cite{DT}\cite{DSlimit}. In two dimensional quantum gravity conformal dimensions are predicted both by lattice models and the corresponding continuum theories and gave the same predictions. It is important to measure these physical quantities. Numerically it has been confirmed that two dimensional quantum space time coupled to $c=-2$ matter show clear fractal scaling and the fractal dimension is consistent with the theoretical value\cite{KKSW}\cite{KW1}. \subsubsection{Microscopic Description of Two Dimensional Random Surface} In general it is highly nontrivial how to relate a microscopic description of lattice theory and the corresponding continuum theory. In two dimensions it is possible to take a continuum limit of lattice theories for some particular cases. It has already been established that the matrix model is powerful tool to solve some lattice models in two dimensions and the procedure how to get the continuum limit is established\cite{DSlimit}. Here we give an example which has well defined microscopic description of $c=0$ lattice model and the continuum limit. We first derive a transfer matrix of two dimensional random surface on the lattice and take the continuum limit and derive the fractal structure of the $c=0$ random surface, a dynamically triangulated lattice without matter\cite{KKMW}. We consider the cylinder amplitude $N(l,l';r;n)$ which counts a number of possible triangulations of random surface for a shape of cylinder with a entrance loop length $l$, a marked exit loop length $l'$, the geodesic distance of the two loops $r$ and the number of triangles $n$. See fig. \ref{fig:composition}. \begin{figure} \begin{center} \begin{minipage}[c]{0.3\textwidth} \epsfxsize=\textwidth \epsfbox{composition.eps} \end{minipage} \end{center} \caption{ Composition law of cylinder amplitude } \label{fig:composition} \end{figure} This quantity satisfies the following composition law: \begin{equation} N(l,l'';r_1+r_2;n_1+n_2) = \sum_{l'}N(l,l';r_1;n_1)N(l',l'';r_2;n_2). \label{eqn:comp-law} \end{equation} We define the following quantity by summing up the triangles with a parameter $K$: \begin{equation} N(l,l',r) \equiv \sum_{n=0}^\infty K^n N(l,l';r;n)\equiv (\hat{N}(r))_{ll'}, \end{equation} which satisfies the following relation \begin{equation} (\hat{N}(r))_{ll'}=((\hat{N}(1))^r)_{ll'}. \end{equation} We can thus claim that $(\hat{N}(1))_{ll'}$ is the transfer matrix of two dimensional surface. We now define the generating function of the transfer matrix by parametrizing $y$ and $y'$ for the entrance and exit loop length, respectively \begin{equation} \hat{N}(y,y';K)\equiv \sum_{l,l'=0}^\infty y^l(y')^{l'}N(l,l';1). \end{equation} It is surprising to recognize that the generating function of the transfer matrix can be obtained by a simple geometric series of the possible one steps \begin{eqnarray} \hat{N}(y,y';K)&=& (2yy'^2K+y^2y'kF)\sum_{n=0}^\infty [yy'2K+y2y'kF+y^2F+yK(F-1)]^n \nonumber \\ &=& \frac{2yy'^2K+y^2y'kF}{1-yy'2K-y2y'kF-y^2F-yK(F-1)}, \label{eqn:trans-gen} \end{eqnarray} where $F$ is a disk amplitude derived by Brezin, Itzykson, Parisi and Zuber\cite{BIPZ}. The term in the parenthesis in the first line of (\ref{eqn:trans-gen}) is possible one step forward with marked point of the exit loop in the dynamical triangulation while the four terms in the square bracket is all the possible steps forward in the entrance loop. See fig.\ref{fig:exit moves} and \ref{fig:entrance moves}. \begin{figure} \begin{center} \begin{minipage}[c]{0.3\textwidth} \epsfxsize=\textwidth \epsfbox{exit.eps} \end{minipage} \caption{ One step exit moves with a marked point with $F$ as disk amplitude } \label{fig:exit moves} \begin{minipage}[c]{0.3\textwidth} \epsfxsize=\textwidth \epsfbox{ent.eps} \end{minipage} \end{center} \caption{ One step entrance moves } \label{fig:entrance moves} \end{figure} It is interesting to note that the geometrical structure of the random surface is directly reflected to the analytic expression of the transfer matrix. For the $c=0$ random surface it is known how to take the continuum limit of random surface for the disk amplitude $F$ by \cite{BIPZ}. Here we simply point out that the continuum limit of the generating function of the transfer matrix can be taken in a well defined way. We can then obtain a continuum expression of the cylinder amplitude by solving differential equation. The details of this interesting example to obtain the continuum expression from the above lattice transfer matrix can be found in \cite{KKMW}. By using the continuum cylinder amplitude we can derive interesting quantities which have direct relation with the fractal structure of the random surface and thus can be measured numerically. Let $\rho(L;R)dL$ be the continuum counterpart of the number of loops belonging to the boundary whose lengths lie between $L$ and $L+dL$. Here $L$ and $R$ are continuum counterparts of $l$ and $r$, respectively. The $\rho(L;R)dL$ turns out to be a scaling function in the continuum limit \begin{eqnarray} \rho(L;R)dL~=~ \Bigl(\frac{3}{7\sqrt{\pi}}\Bigr) \frac{1}{R^2}(x^{-5/2}+\frac{1}{2}x^{-3/2}+\frac{14}{3}x^{1/2})exp(-x) \label{eqn:rho-scale} \end{eqnarray} with $x=L/R^2$ as a scaling parameter. It is very surprising that there is a scaling function for the quantum gravity. This $\rho(L;R)dL$ can be compared with the numerically measured value. \subsubsection{Numerical Results on the Fractal Structure of Two Dimensional Quantum Gravity on the Lattice} It is important to ask what could be the observables of quantum gravity, which could be numerically measured. We have confirmed the fractal nature of the quantum gravity for $c=-2$ model numerically\cite{KKSW}, which is the first numerical confirmation of the fractal nature of quantum gravity. We have analyzed the following scalar fermion models corresponding to $c=-2$ model: \begin{eqnarray} S~&=&~\sum_{T}\int \prod_{i\in T} d \bar{\psi}_i d \psi_i exp\{-\sum_{<ij>}(\bar{\psi}_i - \bar{\psi}_j)(\psi_i - \psi_j)\} \cr &=&~ \sum_{T} det \Delta (T) \cr &=&~ \frac{1}{N+2} T_{N+1} R_{N+1}, \label{eqn:c-2model} \end{eqnarray} where all the possible triangulation $T$ are summed up. The scalar fermions are located on the dynamically triangulated lattice sites and thus $\Delta (T)$ is the lattice laplacian for the triangulation $T$. $T_{N+1}$ and $R_{N+1}$ are the number of tree and rainbow diagrams with $N$ vertices and external lines, respectively. Tree Feynman diagrams of $\phi^3$ theory and planar triangulations have one to one correspondence. Here we fix the topology of the surface as a sphere and the number of triangles to be $N$. $T_n$ and $R_n$ satisfy the same Schwinger-Dyson equation: \begin{eqnarray} T_n~=~\sum_{k=1}^{n-1} T_k T_{n-k}, \end{eqnarray} which leads to the solution $T_n=(2n-2)!/n!(n-1)!$ and $R_n$ has the same form. Since we know the analytic expression of $T_n$ and $R_n$ we can reconstruct arbitrary triangulation of sphere topology by combining the tree and rainbow diagrams with the correct weight. The essential point of the $c=-2$ model is that it has the very simple relation (\ref{eqn:c-2model}) due to the first power of the lattice laplacian. Then $T_n$ and $R_n$ have very simple form and thus numerically huge number of triangulated surfaces can be generated. We have used the formulation called "recursive sampling method" to generate the surface configuration numerically which was initiated by Agistein and Migdal for $c=0$ model \cite{AM}. We have measured several quantities, in particular the number of triangles $V(r)$ within a geodesic distance $r$ parameterized by \begin{equation} V(r)~\equiv~<{\mbox {\rm number of triangles within r steps}}>~ \sim~r^\gamma. \end{equation} We have found that the fractal dimension $\gamma$ approaches a constant value $\gamma \rightarrow 3.55$ with large $N (\rightarrow 5\times 10^6$). This is the clear sign of the fractal scaling of two dimensional random surface\cite{KKSW}. See fig.\ref{fig:frexp1} We have thus numerically confirmed that the fractal structure is the essential observable of two dimensional quantum gravity. In these numerical measurements we needed huge number of triangles to get saturated value of the fractal dimension. Later we have used the formulation of finite size scaling methods applied to the quantum gravity and then we obtained very accurate value of the fractal dimension even with relatively smaller number of triangles but with huge number of configurations\cite{AAIJKWY}. \begin{figure} \begin{center} \begin{minipage}[b]{0.4\textwidth} \epsfxsize=\textwidth \epsfbox{kkmw.eps} \end{minipage} \end{center} \caption{Direct measurement of fractal dimension of $c=-2$ model} \label{fig:frexp1} \end{figure} We define the following quantity: \begin{equation} <L^n(r)> \equiv \sum_{l=1}^\infty l^n \rho(l,r), \end{equation} where $L(r)$ is the length of boundary loops located at $r$ steps from a marked point and $\rho(l,r)$ is the lattice counterpart of the quantity defined in (\ref{eqn:rho-scale}) for a given number of triangles $N$. Finite size scaling formulation of the two dimensional quantum gravity predicts the following formula: \begin{eqnarray} <L^n(r)> &\sim& (N)^{2n/d_F}F_n(y) ~~~~~~ (n\geq 2) \nonumber \\ y&=&\frac{r}{(N)^{1/d_F}}, \end{eqnarray} where $N$ is the number of triangles and $F(y)$ is a scaling function and $y$ is the scaling parameter. As we can see in figs.\ref{fig:finite scaling}, the data are beautifully scaling for $F_2$, $F_3$, $F_4$ with various numbers of triangles which are given in figures. The numerical result for the fractal dimension of quantum space time coupled to $c=-2$ matter by comparing the above formula with the numerical distributions of $F$'s is $$ d_F=3.56 \pm 0.04. $$ \begin{figure} \begin{center} \begin{minipage}[a]{0.3\textwidth} \epsfxsize=\textwidth \epsfbox{copen2.eps} \end{minipage} \begin{minipage}[b]{0.3\textwidth} \epsfxsize=\textwidth \epsfbox{copen3.eps} \end{minipage} \begin{minipage}[c]{0.3\textwidth} \epsfxsize=\textwidth \epsfbox{copen4.eps} \end{minipage} \end{center} \caption{Finite size scaling of $F_2(y)$, $F_3(y)$ and $F_4(y)$ for the number of triangles given in figs.} \label{fig:finite scaling} \end{figure} On the other hand we gave the theoretical prediction of the fractal dimension of two dimensional space time coupled to $c$ matter by Liouville theory \cite{KW1} \begin{eqnarray} d_F&=&2\frac{\sqrt{25-c} + \sqrt{49-c}}{\sqrt{25-c} + \sqrt{1-c}} \nonumber \\ &=& 3.561 ~~~~~~~~ (c=-2). \end{eqnarray} The numerical result and the theoretical data are perfectly consistent. Here we have given an example, $c=-2$ model, of two dimensional quantum gravity which are well understood numerically and analytically. We claim that the essence of the two dimensional quantum gravity is the fractal nature of the space time. \subsection{Susskind Fermion---Staggered Fermion---Dirac K$\ddot{\hbox{a}}$hler Fermion on the Lattice} If we want to put fermions on the lattice, there is the well known chiral fermion problem. The problem states that we cannot put chiral fermions on the lattice with avoiding species doublers\cite{nogo}. Or otherwise it is stated that the anomaly vanishes if we put fermions in a chiral invariant way\cite{Karsten-Smit}. Wilson's proposal on this respect is to introduce non-chiral term in the lattice action and then each species doubler except for one gets a mass of order the inverse lattice constant and thus disappear in the continuum limit\cite{Wilson}. Then the chirality will be recovered and the remaining unique species generates the necessary anomaly in the continuum limit. In fact there is a clear chiral invariant line on the hopping parameter-coupling constant plane where the quak mass and the pion mass vanishes in the Wilson's lattice QCD formulation\cite{Kawamoto}. There is an alternative approach on this problem. The origin of the species doublers are related to the simplicial nature of the manifold. Susskind proposed an idea to collect these species doublers to construct Dirac fermions\cite{Susskind}. For example in four dimensional square lattice there are $2^4=16$ corners of Brillouin zone in the momentum space and each doubler corresponds to a component of the Dirac spinors thus leads to four flavor copies of Dirac fermions. One of the essential points here is to recognize that the species doublers in the configuration space correspond to points located at the center of the simplexes of the fundamental hyper cube. This is related to do with the fact that the Susskind fermion formulation is the flat space version of the Dirac-K$\ddot{\hbox{a}}$hler fermion formulation\cite{DK}\cite{BJ} which is formulated by differential forms and thus independent of the particular choice of lattice shape and thus can be identified as a fermion formulation on the simplicial manifold. Here we explain this situation by starting from the naive fermion formulation. If we naively discretize the Dirac equation on the square lattice we obtain \begin{equation} S_F=\frac{1}{2}\sum_{x,\hat{\mu}}[\overline{\psi}(x)\gamma_\mu\psi (x+\hat{\mu})-\overline{\psi}(x+\hat{\mu})\gamma_\mu\psi(x)], \label{eqn:naive-ferm} \end{equation} where $x$ denotes the lattice site with integer coordinates $x_1,x_2,\cdots,x_d$ in units of lattice spacing which is taken to be 1. $\hat{\mu}$ is the unit vector along the $\mu$th direction. The $\gamma$-matrices satisfy the following Clifford algebra: \begin{equation} \{\gamma_\mu,\gamma_\nu\}=-2\delta_{\mu\nu}, \label{eqn:clifford-alg} \end{equation} with the conventions $\gamma_\mu^\dagger=-\gamma_\mu$ in $d$-dimension. We proposed the following transformation which relates the naive fermion formulation and Susskind fermion formulation \cite{KS}: \begin{equation} \psi(x)=A(x)\chi(x),~~~~~~~~~~~\overline{\psi}(x)= \overline{\chi}(x)A^\dagger(x), \label{eqn:KS-trans} \end{equation} where \begin{equation} A(x)=\gamma_1^{x_1}\gamma_2^{x_2}\cdots\gamma_d^{x_d}. \label{eqn:KS-mat} \end{equation} Then the naive fermion formulation of Dirac action leads to \begin{equation} S_F=-\frac{1}{2}\sum_{x,\hat{\mu},\alpha}\eta_\mu(x) [\overline{\chi}^\alpha(x)\chi^\alpha(x+\hat{\mu})+ \overline{\chi}^\alpha(x+\hat{\mu})\chi^\alpha(x)], \label{eqn:stag-ferm} \end{equation} with \begin{equation} \eta_\mu(x)=(-1)^{x_1+x_2+\cdots+x_{\mu-1}}. \label{eqn:KS-mat} \end{equation} This formulation is called staggered fermion formulation in lattice QCD. The $\gamma$ matrices in the naive fermion formulation have now disappeared in the staggered fermion formulation. Instead we get $C\equiv 2^{[d/2]}$ copies of spinors. We now consider $d$-dimensional hypercube defined on the lattice, with its origin at the site $2y$ and corners \begin{equation} x_\mu=2y_\mu + \eta_\mu, ~~~~~\eta_\mu=0~ \hbox{or}~ 1,~~~ \mu=1,\cdots,d.\nonumber \label{eqn:double-size} \end{equation} Thus $\eta$ labels the set of $C^2$ $d$-dimensional vectors which point on the corners of the hypercube. We introduce the following notations for the convenience: \begin{eqnarray} \chi(2y+\eta)&\equiv& (-1)^y\chi_\eta(y) \nonumber \\ \overline\chi(2y+\eta)&\equiv& (-1)^y\overline\chi_\eta(y) \nonumber \\ (-1)^y&=&(-1)^{y_1+\cdots +y_d}. \end{eqnarray} Then $S_F$ can be written in the following form\cite{Saclay}: \begin{equation} S_F=\sum_{x,\hat{\mu}}\sum_{\eta,\eta'}[\overline{\chi}_\eta(y) \Gamma^\mu_{\eta,\eta'}\Delta_\mu\chi_{\eta'}(y) + \overline{\chi}_\eta(y)\widetilde{\Gamma}^\mu_{\eta,\eta'} \delta_\mu\chi_{\eta'}(y)], \label{eqn:double-space-action} \end{equation} where the first and second lattice derivatives are \begin{eqnarray} \Delta_\mu f(y) &\equiv& \frac{1}{4}[f(y+\hat{\mu})-f(y-\hat{\mu})] \nonumber \\ \delta_\mu f(y) &\equiv& \frac{1}{4}[f(y+\hat{\mu})-2f(y)+f(y-\hat{\mu})]. \end{eqnarray} $\Gamma^\mu$ and $\Gamma^{\mu'}$ are defined by \begin{eqnarray} \Gamma^\mu_{\eta,\eta'}&\equiv&\frac{1}{C}\hbox{tr} (\Gamma_\eta^\dagger\gamma_\mu\Gamma_{\eta'}), \nonumber \\ \widetilde{\Gamma}^\mu_{\eta,\eta'} &\equiv& \frac{1}{C}\hbox{tr} (\Gamma_\eta^\dagger\gamma_\mu\Gamma_{\eta'}) (\delta_{\eta',\eta-\hat{\mu}}-\delta_{\eta',\eta+\hat{\mu}}), \label{eqn:Gamma-def} \end{eqnarray} with \begin{equation} \Gamma_\eta=\gamma_1^{\eta_1}\gamma_2^{\eta_2}\cdots\gamma_d^{\eta_d}. \label{eqn:gamma-prod} \end{equation} Introducing the following notation for the fermionic fields to get rid of the factor of trace in (\ref{eqn:Gamma-def}): \begin{eqnarray} \chi_\eta(y)&=&\sqrt{C}\sum_{\alpha,a}\Gamma^{*\alpha a}_\eta q^{\alpha a}(y) \nonumber \\ \overline{\chi}_\eta(y)&=&\sqrt{C}\sum_{\alpha,a}\overline{q}^{\alpha a}(y) \Gamma^{\alpha a}_\eta, \label{eqn:newferm-def} \end{eqnarray} we obtain the final form of Dirac-K$\ddot{\hbox{a}}$hler fermion action on the flat space \begin{equation} S_F=2^d\sum_{y,\mu}[\overline{q}(y)(\gamma_\mu\otimes{\bf 1}) \Delta_\mu q(y) + \overline{q}(y)(\gamma^\dagger_5\otimes t_\mu^\dagger t_5^\dagger) \delta_\mu q(y)]. \end{equation} The fermionic part of this action can be interpreted as Susskind fermions while this action has bosonic counterparts as well. The bosonic part including second derivative is higher order than the fermionic part with respect to the lattice constant. In the fermion bilinears the first matrix acts on Greek indices interpreted as Dirac indices, the second one on Latin indices interpreted as flavor indices. $t_\mu$ stands for the matrix $\gamma^*_\mu$ when acting on flavor space, and $\gamma_5=\gamma_1\gamma_2\cdots \gamma_d$. \subsection{ Chern-Simons Gravity and Ponzano-Regge Model} \subsubsection{Chern-Simons Gravity} In this subsection we explain formulations of three dimensional gravity. We first summarize the Chern-Simons gravity formulated by Witten\cite{Witten1}. We choose the gauge group as Euclidean version of three dimensional Poincare group $ISO(3)$. Then we define one form gauge field and zero form gauge parameter as \renewcommand{\arraystretch}{1.5} \begin{equation} \begin{array}{rcl} A_\mu&=&e^a_\mu P_a + \omega^a_\mu J_a, \\ v &=& \rho^a P_a + \tau^a J_a, \end{array} \end{equation} where $e^a_\mu$ and $\omega^a_\mu$ are dreibein and spin connection, respectively, and $\rho$ and $\tau$ are the corresponding gauge parameters. The momentum generator $P_a$ and the angular momentum generator $J_a$ of $ISO(3)$ satisfy \begin{equation} [J_a,J_b] = \epsilon_{abc} J^c, ~~~ [J_a,P_b] = \epsilon_{abc} P^c, ~~~ [P_a,P_b] = 0. \end{equation} Using the invariant quadratic form which is special in three dimensions, we can define the inner product \begin{equation} \langle J_a,P_b \rangle = \delta_{ab}, ~~\langle J_a,J_b \rangle = \langle P_a,P_b \rangle = 0. \end{equation} We then obtain Einstein-Hilbert action of three dimensional gravity from Chern-Simons action \begin{equation} \int \Bigl\langle AdA + \frac{2}{3}A^3 \Bigr\rangle = \int \epsilon^{\mu\nu\rho}e_{\mu a}F^a_{\nu\rho}~d^3x, \end{equation} where \begin{equation} F^a_{\mu\nu}= \partial_\mu\omega^a_\nu - \partial_\nu\omega^a_\mu + \epsilon^a_{~bc}\omega^b_\mu\omega^c_\nu.\label{CSCurvature} \end{equation} The component wise gauge transformation of $\delta A_\mu = - D_\mu v$ is given by \begin{equation} \begin{array}{rcl} \delta e^a_\mu &=& -D_\mu\rho^a -\epsilon^{abc}e_{\mu b}\tau_c,\\ \delta \omega^a_\mu &=& -D_\mu\tau^a.\label{CS gauge transformation} \end{array} \end{equation} At this stage it is important to recognize that the local Lorentz transformation is generated by the gauge parameter $\tau$ \begin{equation} \begin{array}{rcl} \delta e^a_\mu &=& -\epsilon^{abc}e_{\mu b}\tau_c,\\ \delta \omega^a_\mu &=& -D_\mu\tau^a, \label{LLGT} \end{array} \end{equation} while the gauge transformation of diffeomorphism is generated by the gauge parameter $\rho$ \begin{equation} \begin{array}{rcl} \delta e^a_\mu &=& -D_\mu\rho^a, \\ \delta \omega^a_\mu &=& 0.\label{DGT} \end{array} \end{equation} Three dimensional Einstein gravity is thus elegantly formulated by Chern-Simons action. This is essentially related to the fact that the three dimensional Einstein gravity does not include dynamical graviton and thus can be formulated by the topological Chern-Simons action. The equivalence of the above action and Einstein- Hilbert action is, however, valid only if the dreibein $e^a_\mu$ is invertible. The quantization and perturbative renormalizability around the nonphysical classical background $e^a_\mu = 0$ is the natural consequence of the formulation. \subsubsection{Ponzano-Regge Model} Ponzano and Regge noticed that angular momenta of 6-$j$ symbol can be identified as link lengths of a tetrahedron\cite{Ponzano-Regge}. \begin{figure} \begin{center} \begin{minipage}[b]{0.4\textwidth} \epsfxsize=\textwidth \epsfbox{tetra.eps} \end{minipage} \end{center} \caption{tetrahedron with angular momenta on the links} \label{fig:colored tetra} \end{figure} In particular they showed the following approximate relation: \begin{equation} (-1)^{\sum_{i=1}^{6} J_{i}} \sixj{J_1}{J_2}{J_3}{J_4}{J_5}{J_6} \sim \frac{1}{\sqrt{12 \pi V}} \cos \left( S_{\hbox{{\scriptsize Regge}}} + \frac{\pi}{4} \right) \quad (\hbox{all } J_i \gg 1) , \label{P-R formula} \end{equation} where $S_{\hbox{{\scriptsize Regge}}}$ is the Regge action of Regge calculus\cite{Regge} for a tetrahedron having link length $J_k$ $(k=1\sim 6)$ which correspond to the angular momentum of the corresponding 6-$j$ symbol and $V$ is the volume of the tetrahedron. Based on this observation they proposed the following partition function: \begin{equation} Z_{\hbox{{\scriptsize PR}}} = \lim_{\lambda \rightarrow \infty} \sum_{J \leq \lambda} \prod_{\hbox{\tiny vertices}} \Lambda (\lambda) ^{-1} \prod_{\hbox{\tiny edges}} (2J+1) \prod_{\hbox{\tiny tetrahedra}} (-1)^{\sum J_i} \sixj{J_1}{J_2}{J_3}{J_4}{J_5}{J_6} \label{ZofPR}. \end{equation} Thus the partition function $Z_{\hbox{{\scriptsize PR}}}$ is the product of the partition function of each tetrahedron which reproduces the cosine of the Regge action in contrast with the exponential of the Regge action in Regge calculus. There is an argument about the origin of the cosine, that right and left handed contributions of the general coordinate frames contribute separately and thus the summation of the exponential with the different sign factor for the Regge action appears. It is thus natural to expect that this action leads to a gravity action. Important characteristic of the Ponzano-Regge action is that it has a topological nature on a simplicial manifold. The action is invariant under the following 2-3 and 1-4 Alexander moves. The 2-3 and 1-4 moves are related to the following 6-$j$ relations: \begin{eqnarray} && \sum_{K} (-1)^{K+\sum_{i=1}^{9}J_i} (2K+1) \sixj{J_1}{J_8}{K}{J_7}{J_2}{J_3} \sixj{J_7}{J_2}{K}{J_6}{J_9}{J_4} \sixj{J_6}{J_9}{K}{J_8}{J_1}{J_5} \nonumber \\ &=& \sixj{J_3}{J_4}{J_5}{J_6}{J_1}{J_2} \sixj{J_3}{J_4}{J_5}{J_9}{J_8}{J_7}, \label{2-3move} \end{eqnarray} and \begin{eqnarray} && \sum_{K_i} \left[ \prod_{i=1}^{4} (2K_i + 1) \right] (-1)^{\sum K_i} \Lambda (\lambda)^{-1} \sixj{J_1}{J_2}{J_3}{K_1}{K_2}{K_3} \sixj{J_4}{J_6}{J_2}{K_3}{K_1}{K_4} \nonumber \\ && \hspace{2cm} \times \sixj{J_3}{J_4}{J_5}{K_4}{K_2}{K_1} \sixj{J_1}{J_5}{J_6}{K_4}{K_3}{K_2} = (-1)^{\sum J_i} \sixj{J_1}{J_2}{J_3}{J_4}{J_5}{J_6}. \label{1-4move} \end{eqnarray} The geometrical correspondence of 2-3 and 1-4 moves with two tetrahedra into three tetrahedra and one tetrahedron into four tetrahedra is obvious from fig.\ref{fig:Alexander move}. \begin{figure} \begin{center} \begin{minipage}[b]{0.4\textwidth} \epsfxsize=\textwidth \epsfbox{Alexander2.eps} \end{minipage} \hspace*{2cm} \begin{minipage}[t]{0.4\textwidth} \epsfxsize=\textwidth \epsfbox{Alexander1.eps} \end{minipage} \end{center} \caption{2-3 move and 1-4 move} \label{fig:Alexander move} \end{figure} In the formula of 1-4 move there appears the following infinite sum which is then introduced as a regularization factor in the denominator with a cutoff $\lambda$: \begin{eqnarray} \Lambda (\lambda) &=& \frac{1}{2 J_1 + 1} \!\!\!\!\!\!\!\!\!\! \sum_{ \mbox{ \scriptsize $ \begin{array}{c} K_2,K_3 \leq \lambda,\\ |K_2 - K_3| \leq J_1 \leq K_2 + K_3 \end{array} $} } \!\!\!\!\!\!\!\!\!\! (2 K_2 + 1)(2 K_3 + 1) \nonumber \\ &=& \sum_{J=0}^{\lambda} (2J+1)^2 \sim \frac{4 \lambda^3}{3} ~~ (\lambda \rightarrow \infty) \label{Lambda}. \end{eqnarray} It is known that these two Alexander moves reproduce any three dimensional simplicial manifold. Thus the partition function $Z_{\hbox{{\scriptsize PR}}}$ is invariant under the variation of metric and is expected to be topological. There is a beautiful generalization of this model to avoid the above cut off dependence by introducing the quantum group formulation as a regularization by Turaev and Viro\cite{Turaev-Viro}. Their proposal triggered number of later investigations for three dimensional lattice gravity\cite{Ooguri-Sasakura}\cite{Boulatov}\cite{3dgravity}. Turaev, Ooguri-Sasakura and Boulatov suggested the equivalence of the Chern-Simons gravity and the Ponzano-Regge model\cite{Turaev-Viro}\cite{Ooguri-Sasakura}\cite{Boulatov}. In the later section we show that the continuum limit of the lattice Ponzano-Regge model leads to the Chern-Simons gravity by explicitly constructing a lattice gauge gravity model. \subsection{Four Dimensional Gravity on the Lattice } In analogy with the formulation of three dimensional gravity by Chern-Simons action and 6-$j$ symbols of the Ponzano-Regge model, it is possible to formulate a four dimensional topological gravity by $BF$ theory and 15-$j$ symbols \cite{Ooguri}. There are variants of the similar four dimensional models developed later \cite{FDgravity}\cite{Smolin}. Ooguri proposed the following model on the four dimensional simplicial manifold composed of four dimensional fundamental simplexes: \begin{equation} Z=\sum_C\frac{1}{N_{\hbox{\tiny sym}}}\lambda^{N_4(C)}Z_C, \end{equation} where the summation $\sum_C$ is over oriented four dimensional symplicial manifolds, and $N_{\hbox{\tiny sym}}$ is a symmetric factor and $N_4(C)$ is a number of 4-simplexes in $C$. The partition function for a given $C$ is \begin{eqnarray} Z_C&=&\sum_J\prod_{t:\hbox{\tiny triangles}}(2J_t+1)\prod_{\hbox{\tiny tetrahedra}} \{6j\}\prod_{\hbox{\tiny 4-simplexes}}\{15j\} \\ &=&\sum_J\sum_{m,n}\prod_{t:\hbox{triangles}}(2J_t+1) \prod_{T:\hbox{\tiny tetrahedra}} \int dU_T D^{J_1,T}(U_T)\cdots D^{J_4,T}(U_T), \label{eqn:15j} \end{eqnarray} where the first line is the symbolic presentation of the geometrical nature of the four dimensional fundamental simplex which has ten triangles, each of which corresponds to 3-$j$ symbol, and ten 3-$j$ symbols lead 15-$j$ symbol, which is analogous to the situation where 6 angular momenta were assigned to the links of tetrahedron and four triangles of a tetrahedron carry four 3-$j$ symbols which leads to a 6-$j$ symbol in three dimensions. By using the orthonormality of the 6-$j$ symbols one can show that the results of this summation is independent of the choice of splitting of the vertices and depends only on the structure of the complex $C$, where extra 6-$j$ symbols are necessary to keep this invariance. This is again analogous to the Alexander move invariance of the Ponzano-Regge model. The second line of eq.(\ref{eqn:15j}) comes from the fact that the 15-$j$ and 6-$j$ symbols are written down by 3-$j$ symbols and the products of 3-$j$ symbols can be replaced by the products of $D$ functions by using the following relations: \begin{eqnarray} \int DU D^{J_1}_{m_1n_1}(U) D^{J_2}_{m_2n_2}(U&)& D^{J_3}_{m_3n_3}(U) = \threej{J_1}{J_2}{J_3}{m_1}{m_2}{m_3}\threej{J_1}{J_2}{J_3}{n_1}{n_2}{n_3}, \label{eqn:3d3j}\\ \sixj{J_1}{J_2}{J_3}{J_4}{J_5}{J_6} = \sum_{\hbox{\tiny{all $m_i$}}} (-&1&)^{\sum_i (J_i - m_i)} \threej{J_1}{J_2}{J_3}{-m_1}{-m_2}{-m_3}\nonumber \\ &&\hspace{-2cm} \times \threej{J_1}{J_5}{J_6}{m_1}{-m_5}{m_6} \threej{J_4}{J_2}{J_6}{m_4}{m_2}{-m_6} \threej{J_4}{J_5}{J_3}{-m_4}{m_5}{m_3}, \label{eqn:6j3j} \end{eqnarray} where the 3-$j$ symbols carry $m_i$ suffices which correspond to the third components of the angular momentum $J_i$. Each tetrahedron $T$ carries a group element $U_T$, and $J_{1,T},\cdots J_{4,T}$ are the spins on the links dual to the four triangles of $T$. The matrix elements $D^{J}$'s are multiplied around triangles in $C$. Each triangle $t$ is shared by a finite number of tetrahedra; $T_1{(t)},T_2{(t)},\cdots,T_{n_t}{(t)}$. We can perform the sum over the spin-$J_t$ on the link dual to $t$ using the formula \begin{equation} \sum_{J_t}(2J_t+1)\hbox{Tr}\left[ D^{J_t}(U_{T_1}^{(t)}) \cdots D^{J_t}(U_{T_{n_t}}^{(t)})\right] = \delta(U_{T_1}^{(t)}\cdots U_{T_{n_t}}^{(t)},1). \label{eqn:flat-con} \end{equation} After carrying out $dU$ integration we obtain the condition that the holonomy around the triangle $t$ is trivial, the lattice version of flat connection condition. The above result suggests that $Z_C$ is related to the partition function of the $BF$ model which is defined for an oriented manifold $M$ as \begin{equation} Z_{BF} = \int DB~DA \hbox{exp}(i\int_M <B,dA+[A,A]>), \end{equation} where $A$ is a connection one-form on the manifold $M$, while $B$ is a Lie algebra valued two-form and $<,>$ is the invariant bilinear form on the Lie algebra. In this so called $BF$ action\cite{BF} with $F\equiv dA+[A,A]$, we can consider $B$ as Lagrange multiplier then we obtain the flat connection condition which is equivalent to the lattice version obtained from eq.(\ref{eqn:flat-con}). This action is the four dimensional counterpart of the Chern-Simons action in the sense that the flat connection condition is the equation of motion. We can thus naively expect that the 15-$j$ model leads to the $BF$ theory in the continuum limit. This is similar to the situation that the Ponzano-Regge model leads to the Chern-Simons gravity in the continuum limit which we are going to show in the later section. \setcounter{equation}{0} \section{A Possible Formulation towards Gauge Gravity coupled to Matter on the Simplicial Lattice Manifold} \subsection{A Possible Formulation towards Unified Model on the Lattice} From the known results which we have collected in the last subsections, I intend to figure out my personal view on what could be the necessary formulation to realize the possible formulations towards unified model on the lattice. I found that the example of Ising spin on the square lattice and dynamically triangulated lattice are the most convincing examples of the following picture: a regularized gravitational background can be generated by the dynamical triangulation and the matter field fermions can be generated by some degrees of freedom of fields located on the simplexes of a simplicial manifold. Lattice models of conformal field theories give many more varieties of the similar examples where matter central charge $c$ takes different values for different type of lattice models. One may wonder it would be special case of two dimensional field theory due to the conformal invariance on the second order phase transition point of two dimensional lattice models. I find it is very natural to expect to have the similar mechanism working even in higher dimensions. It would be, however, very difficult to solve four dimensional lattice models analytically while it was possible to solve two dimensional model analytically and derive the corresponding conformal invariant continuum models. As we have seen in the analyses of the fractal structure of quantum gravity, numerical simulation played an important role. In the two dimensional example of $c=-2$ model theoretical and numerical analyses on the fractal dimension were perfectly consistent. Since we don't expect the solvability in higher dimensional lattice gravity models with matter, numerical simulations may play an important role again. In considering the present situation, the lattice QCD would be a good example to find analogy in the lattice gravity formulation. Lattice QCD is not solvable but is perfectly correct formulation of QCD even in non perturbative regime. The Lorentz invariance is lost but the gauge invariance was strictly kept in the finite lattice regime. We need to find a gravitational counterpart of the lattice QCD formulation. In three dimensions Chern-Simons gravity is formulated as a gauge theory. It would be thus very instructive to find lattice gravity formulation of Chern-Simons gravity. As we show in the next section it is possible to formulate the Chern-Simons gravity in analogy with Wilson's lattice QCD formulation. What is surprising there is that the continuum limit of the Chern-Simons gravity can be considered analytically with the help of Ponzano-Regge model. We find it is instructive that the three dimensional gravity is formulated by gauge theory of one form with Chern-Simons action. If one can formulate a gauge theory in terms of forms, the general coordinate invariance is automatic. On the other hand it is also very natural to expect that the gravity theory can very well be formulated by a gauge theory. If we intend to formulate a lattice gravity theory as a gauge theory we need to formulate gauge theory in terms of differential forms. Furthermore differential forms have very natural correspondence with simplexes on the simplicial manifold. As we can see in fig.\ref{fig:glatgr} the fields of differential forms are naturally put on the simplexes. \begin{figure} \begin{center} \begin{minipage}[b]{0.4\textwidth} \epsfxsize=\textwidth \epsfbox{glatgr.eps} \end{minipage} \end{center} \caption{Differential forms on the symplices} \label{fig:glatgr} \end{figure} We thus propose that the generalized gauge theory which are formulated by differential forms would be a good candidate to formulate lattice gravity theory even with matter fermions, possibly Dirac-K$\ddot{\hbox{a}}$hler fermions, which are also formulated by differential forms\cite{DK}\cite{BJ}. \subsection{Generalized Gauge Theory} The generalized Chern-Simons actions, which were proposed by the present author and Watabiki about ten years ago, is a generalization of the ordinary three-dimensional Chern-Simons theory into arbitrary dimensions~\cite{KW2}. We summarize the results in this section. The essential point of the generalization is to extend a one-form gauge field and zero-form gauge parameter to a quaternion valued generalized gauge field and gauge parameter which contain forms of all possible degrees. Correspondingly the standard gauge symmetry is extended to much higher topological symmetry. These generalizations are formulated in such a way that the generalized actions have the same algebraic structure as the ordinary three-dimensional Chern-Simons action. Since this generalized Chern-Simons action is formulated completely parallel to the ordinary gauge theory, the generalization can be extended further to the topological Yang-Mills action and generalized Yang-Mills actions of arbitrary dimensions. In the most general form, a generalized gauge field $\cal{A}$ and a gauge parameter $\cal{V}$ are defined by the following component form: \begin{eqnarray} {\cal A} & = & {\mbox{\bf 1}}\psi + {\mbox{\bf i}} \hat{\psi} + {\mbox{\bf j}} A + {\mbox{\bf k}} \hat{A}, \label{eqn:ggf} \\ {\cal V} & = & {\mbox{\bf 1}} \hat{a} + {\mbox{\bf i}} a + {\mbox{\bf j}} \hat{\alpha} + {\mbox{\bf k}} \alpha, \label{eqn:ggp} \end{eqnarray} where $( \psi, \alpha )$, $( \hat{\psi}, \hat{\alpha} )$, $( A, a )$ and $( \hat{A},\hat{a} )$ are direct sums of fermionic odd forms, fermionic even forms, bosonic odd forms and bosonic even forms, respectively, and they take values on a gauge algebra. The bold face symbols ${\bf 1}$, ${\bf i}$, ${\bf j}$ and ${\bf k}$ satisfy the algebra \begin{equation} \begin{array}{c} {\bf 1}^2={\bf 1}, \quad {\bf i}^2=\epsilon_1 {\bf 1}, \quad {\bf j}^2=\epsilon_2 {\bf 1}, \quad {\bf k}^2=-\epsilon_1 \epsilon_2 {\bf 1}, \\ {\bf i}{\bf j}=-{\bf j}{\bf i}={\bf k}, \quad {\bf j}{\bf k}=-{\bf k}{\bf j}=-\epsilon_2 {\bf i}, \quad {\bf k}{\bf i}=-{\bf i}{\bf k}=-\epsilon_1 {\bf j}, \end{array} \end{equation} where $(\epsilon_1,\epsilon_2)$ takes the value $(-1,-1), (-1,+1), (+1,-1)$ or $(+1,+1)$. Throughout this paper we adopt the convention $(\epsilon_1,\epsilon_2)=(-1,-1)$ unless otherwise stated, then the above algebra corresponds to the quaternion algebra. The following graded Lie algebra can be adopted as a gauge algebra: \begin{eqnarray} \left[ T_a , T_b \right] &=& f^c_{ab} T_c, \nonumber\\ \left[ T_a , \Sigma_\beta \right] &=& g^\gamma_{a\beta} \Sigma_\gamma, \\ \left\{ \Sigma_\alpha , \Sigma_\beta \right\} &=& h^c_{\alpha\beta} T_c, \nonumber \end{eqnarray} where all the structure constants are subject to consistency conditions which follow from the graded Jacobi identities. If we choose $\Sigma_\alpha=T_a$ especially, this algebra reduces to $T_a T_b = k^c_{ab}T_c$ which is closed under multiplication. A specific example of such algebra is realized by Clifford algebra~\cite{KW3}. The components of the gauge field $\cal{A}$ and parameter $\cal{V}$ are assigned to the elements of the gauge algebra in a specific way: \begin{equation} \begin{array}{rclrclrclrcl} A &=& T_a A^a, \quad \hat{\psi} &=& T_a \hat{\psi}^a, \quad \psi &=& \Sigma_\alpha \psi^\alpha, \quad \hat{A} &=& \Sigma_\alpha \hat{A}^\alpha, \\ \hat{a} &=& T_a \hat{a}^a, \quad \alpha &=& T_a \alpha^a, \quad \hat{\alpha}&=&\Sigma_\alpha \hat{\alpha}^\alpha, \quad a &=& \Sigma_\alpha a^\alpha. \end{array} \end{equation} An element having the same type of component expansion as $\cal{A}$ and $\cal{V}$ belong to $\Lambda_-$ and $\Lambda_+$ class, respectively, and these elements fulfill the following $Z_2$ grading structure: $$ [\lambda_+ , \lambda_+] \in \Lambda_+, \qquad [\lambda_+ , \lambda_-] \in \Lambda_-, \qquad \{\lambda_- , \lambda_-\} \in \Lambda_+, $$ where $\lambda_+ \in \Lambda_+$ and $\lambda_- \in \Lambda_-$. The elements of $\Lambda_-$ and $\Lambda_+$ can be regarded as generalizations of odd forms and even forms, respectively. In particular the generalized exterior derivative which belongs to $\Lambda_-$ is given by \begin{equation} Q={\bf j}d, \end{equation} and the following relations similar to the ordinary differential algebra hold: $$ \{Q , \lambda_-\}=Q\lambda_-, \qquad [Q , \lambda_+]=Q\lambda_+, \qquad Q^2=0, $$ where $\lambda_+ \in \Lambda_+$ and $\lambda_- \in \Lambda_-$. To construct the generalized Chern-Simons actions, we need to introduce two kinds of traces \begin{equation} \begin{array}{rcl} \mbox{Tr}[T_a , \cdots]=0, &{\qquad}& \mbox{Tr}[\Sigma_\alpha , \cdots]=0, \\ \mbox{Str}[T_a , \cdots]=0, &{\qquad}& \mbox{Str}\{\Sigma_\alpha , \cdots\}=0, \\ \end{array} \label{eqn:gtr} \end{equation} where $(\cdots)$ in the commutators or the anticommutators denotes a product of generators. In particular $(\cdots)$ should include an odd number of $\Sigma_\alpha$'s in the last eq. of (\ref{eqn:gtr}). $\mbox{Tr}$ is the usual trace while $\mbox{Str}$ is the super trace satisfying the above relations. These definitions of the traces are crucial to show that the generalized gauge theory action can be invariant under the generalized gauge transformation presented bellow. As we have seen in the above the generalized gauge field ${\cal A}$ and parameter ${\cal V}$, and the generalized differential operator $Q$ play the same role as the one-form gauge field and zero-form gauge parameter and differential operator of the usual gauge theory, respectively. We can then construct generalized actions in terms of these generalized quantities. We first define a generalized curvature \begin{equation} {\cal F} \equiv \{Q+{\cal A},Q+{\cal A}\} = Q{\cal A} + {\cal A}^2. \label{eqn:gcurv} \end{equation} We can construct generalized actions of Chern-Simons form, topological Yang-Mills form and Yang-Mills form. They have the standard forms with respect to the generalized quantities \begin{eqnarray} S_{GCS}&=&\int_M \mbox{Tr}^* \left( \frac{1}{2}{\cal A}Q{\cal A}+\frac{1}{3}{\cal A}^3 \right), \nonumber \\ S_{GYM}&=&\int_M \mbox{Tr}^* \left( {\cal F} {\cal F} \right), \label{eqn:GAS} \\ S_{GYM}&=&\int_M \mbox{Tr}^* \left( {\cal F} *{\cal F} \right),\nonumber \end{eqnarray} where $*$ is generalized Hodge star defined later. The product of the generalized fields should be understood as wedge product. The generalized trace $\mbox{Tr}^*$ have several types: $\mbox{Tr}^*=\mbox{Tr}_{\bf q}$ or $\mbox{Tr}^*=\mbox{Str}_{\bf q}$. $\mbox{Tr}_{\bf q}(\cdots)$ and $\mbox{Str}_{\bf q}(\cdots) \ \ ({\bf q}={\bf 1}, {\bf i}, {\bf j}, {\bf k})$ are defined so as to pick up only the coefficients of ${\bf q}$ from $(\cdots)$ and take the traces defined by eq.(\ref{eqn:gtr}). The reason why we obtain the four different types of action is related to the fact that the generalized actions in the trace are quaternion valued and thus we should pick up particular combination in order to keep the generalized gauge invariance. For example the generalized Chern-Simons action $\frac{1}{2}{\cal A}Q{\cal A}+\frac{1}{3}{\cal A}^3 $, belongs to $\Lambda_-$ class and thus possesses the four different component types, the same types as in ${\cal A}$ of (\ref{eqn:ggf}). For example the ${\bf k}$-th component of the generalized Chern-Simons action is even dimensional bosonic component and thus the action \begin{equation} \displaystyle{S^e_{GCS}= \int_M \mbox{Tr}_{\bf k} \left( \frac{1}{2}{\cal A}Q{\cal A}+\frac{1}{3}{\cal A}^3 \right)} \label{eqn:egcs} \end{equation} is even dimensional bosonic action while ${\bf j}$-th component of the generalized Chern-Simons action is odd dimensional bosonic component and thus the action \begin{equation} \displaystyle{S^o_{GCS}=\int_M \mbox{Str}_{\bf j} \left( \frac{1}{2}{\cal A}Q{\cal A}+\frac{1}{3}{\cal A}^3 \right)} \label{eqn:ogcs} \end{equation} is odd dimensional bosonic action. It is also possible to obtain fermionic generalized Chern-Simons actions by taking ${\bf 1}$-th and ${\bf i}$-th components. We then need to pick up $d$-form terms to obtain $d$ dimensional actions defined on a $d$-dimensional manifold $M$. We can thus construct any dimensional generalized Chern-Simons actions. On the other hand the generalized topological Yang-Mills action ${\cal F} {\cal F}$ belongs to $\Lambda_+$ class and thus possesses the four different component types, the same types as in ${\cal V}$ of (\ref{eqn:ggp}). For example the ${\bf 1}$-th component of the generalized topological Yang-Mills action is even dimensional bosonic component and thus the action \begin{equation} \displaystyle{S_{GYM}=\int_M \mbox{Str}_{\bf 1} \left( {\cal F} {\cal F} \right)} \label{eqn:gyma} \end{equation} is even dimensional bosonic action. We can similarly construct odd dimensional bosonic action by taking ${\bf i}$-th component of the generalized topological Yang-Mills action. Important fact is that these generalized Chern-Simons actions and the topological Yang-Mills actions are invariant under the following generalized gauge transformations: \begin{equation} \delta {\cal A}=[Q+{\cal A} , {\cal V} ], \label{eqn:ggt} \end{equation} where ${\cal V}$ is the generalized gauge parameter defined by eq.(\ref{eqn:ggp}). It should be noted that this symmetry is much larger than the usual gauge symmetry, in fact topological symmetry, since the gauge parameter ${\cal V}$ contains as many gauge parameters as gauge fields of various forms in ${\cal A}$. There is another suggestive topological nature due to the parallel construction with the standard gauge theory. In ordinary gauge theory the integral for the $n$-th power of the trace of curvature is called Chern character and has topological nature. In the generalized gauge theory it is possible to define generalized Chern character which is expected to have topological nature related to \lq\lq generalized index theorem" \begin{eqnarray} {\mbox{Str}}_{\bf\mbox{1}}({\cal{F}}^n) &=&{\mbox{Str}}_{\bf\mbox{1}}(Q\Omega_{2n-1}), \label{eqn:bepo} \\ {\mbox{Tr}}_{\bf\mbox{i}}({\cal{F}}^n) &=&{\mbox{Tr}}_{\bf\mbox{i}}(Q\Omega_{2n-1}), \label{eqn:bopo} \end{eqnarray} where $\Omega_{2n-1}$ is the \lq\lq generalized" Chern-Simons forms. Especially, for $n=2$ case in (\ref{eqn:bepo}), we obtain the topological Yang-Mills type action of (\ref{eqn:gyma}) related to the generalized Chern-Simons action with one dimension lower on an even-dimensional manifold $M$, \begin{equation} \int_M{\mbox{Str}}_{\bf\mbox{1}}{\cal F}^2 =\int_M{\mbox{Str}}_{\bf\mbox{1}}\Bigg( Q\Big({\cal{A}}Q{\cal{A}} +\frac{2}{3}{\cal A}^3\Big)\Bigg), \label{eqn:2cc} \end{equation} which has the same form of the standard relation. In the case of the generalized Yang-Mills action the story is different. In order to define the generalized Yang-Mills action we need to define the dual of the generalized curvature by defining Hodge star operation which breaks the gauge symmetry of the higher forms as will be explained later. Correspondingly the generalized Yang-Mills action is not invariant under the generalized gauge transformation (\ref{eqn:ggt}). If we, however, restrict to use the zero form gauge parameter then the action recovers gauge invariance. It is, however, important to recognize that the generalized Yang-Mills formulation is essentially the general realization of the non-commutative geometry formulation of gauge theory \`{a} la Connes\cite{Connes}. We can thus formulate the Weinberg-Salam model by our generalized Yang-Mills action which will be explained later. \setcounter{equation}{0} \section{Gravity on the Lattice} \subsection{First Step towards the Generalized Chern-Simons Actions on the Lattice } \label{First Step} Here we first show the concrete expressions of generalized Chern-Simons actions in two, three and four dimensions. For generalized gauge fields and parameters we introduce the following notations: \begin{eqnarray} {\cal A} & = & {\mbox{\bf j}} A + {\mbox{\bf k}} \hat{A} \nonumber \\ &\equiv& {\mbox{\bf j}}(\omega + \Omega) + {\mbox{\bf k}}( \phi + B + \epsilon_1 H) \label{eqn:def-gf} \\ {\cal V} & = & {\mbox{\bf 1}} \hat{a} + {\mbox{\bf i}} a \nonumber \\ &\equiv& {\mbox{\bf 1}}(v + \epsilon_1 b + V) + {\mbox{\bf i}}( u + U), \label{eqn:def-gp} \end{eqnarray} where we have omitted the fermionic gauge fields $\psi, \hat{\psi}$ and gauge parameters $\alpha, \hat{\alpha}$ for simplicity. Here zero, one, two, three, four form gauge fields and gauge parameters are denoted as $\phi, \omega, B, \Omega, H$ and $v, u, b, U, V$, respectively. By substituting eq.(\ref{eqn:def-gf}) into $S^e_{GCS}$ and $S^o_{GCS}$ of eq.(\ref{eqn:egcs}) and eq.(\ref{eqn:ogcs}) and picking up two, three and four form part of the action, we can obtain more explicit expressions of our generalized Chern-Simons actions in two, three and four dimensions. \begin{eqnarray} S_2 &=& - \int \mbox{Tr}\{\phi(d\omega + \omega^2) + \phi^2 B\}, \label{S_2} \\ S_3 &=& - \int \mbox{Str}\{\frac{1}{2}\omega d\omega + \frac{1}{3}\omega^3 - \phi(d B + [\omega,B]) + \phi^2 \Omega \},\label{S_3} \\ S_4 &=& - \int \mbox{Tr}\{ B(d\omega + \omega^2) + \phi(d \Omega + \{\omega, \Omega\}) + \phi B^2 + \phi^2 H \}. \label{S_4} \end{eqnarray} These actions are invariant under the following gauge transformations: \begin{eqnarray} \delta \phi &=& [\phi,v], \nonumber \\ \delta \omega &=& dv + [\omega,v] - \{\phi,u\}, \nonumber \\ \delta B &=& du + \{\omega,u\} + [B,v] + [\phi,b], \nonumber \\ \delta \Omega &=& db + [\omega,b] + [\Omega,v] - \{B,u\} + \{\phi,U\},\nonumber \\ \delta H &=& -dU-\{\omega,U\}+\{\Omega,u\}+[H,v]+[B,b]+[\phi,V], \label{eqn:com-ggt} \end{eqnarray} which are obtained by substituting eqs.(\ref{eqn:def-gf}) and (\ref{eqn:def-gp})into the gneralized gauge transformations (\ref{eqn:ggt}). Equations of motion derived from the actions $S_2$, $S_3$ and $S_4$ are \begin{eqnarray} \phi^2 \ &=& \ 0 , \nonumber \\ d \phi \, + \, [ \, \omega , \phi \, ] \ &=& \ 0 , \nonumber \\ d \omega \, + \, \omega^2 \, + \, \{ \phi , B \} \ &=& \ 0 , \nonumber \\ d B \, + \, [ \, \omega , B \, ] \, + \, [ \, \Omega , \phi \, ] \ &=& \ 0 , \nonumber \\ d \Omega \, + \, \{ \omega , \Omega \} \, + \, B^2 \, + \, \{ \phi , H \} \ &=& \ 0 , \label{eqn:com-eqmotions} \end{eqnarray} which are component-wise expressions of the vanishing generalized curvature ${\cal F}=0$ as equations of motion of the generalized Chern-Simons action, obtained by substituting (\ref{eqn:def-gf}) into (\ref{eqn:gcurv}). Classically we have already shown that gravity theories can be formulated by using the generalized Chern-Simons actions of two, three and four dimensions. In two dimensions $SL(2,R)$ topological gravity formulated by Verlinde and Verlinde\cite{Verlinde-Verlinde} can be formulated as a local gauge theory of the generalized Chern-Simons action by using Clifford algebra\cite{KW3}. In three dimensions the gravity theory formulated by generalized Chern-Simons action with three dimensional Super Poincar\`{e} algebra\cite{Yoshida} leads to the standard Chern-Simons gravity of Einstein-Hilbert action formulated by Witten\cite{Witten1}. Thus classically three dimensional Einstein gravity can be formulated by both the standard Chern-Simons action and the generalized Chern-Simons action. In four dimensions a topological conformal gravity can be formulated by the generalized Chern-Simons action with Conformal algebra\cite{KW3}. In these analyses the zero form gauge field, as a classical solution of the equation motion, played a role of \lq\lq order parameter" of gravity phase. In other words the above mentioned gravity theories are realized when the zero form part of the classical solution vanishes. Classically gravity theories in two, three and four dimensions are formulated by the generalized Chern-Simons actions. We have recently shown that the quantization of the generalized Chern-Simons actions can be carried out\cite{KSTU}. It turned out that the quantization of the generalized Chern-Simons action is highly nontrivial due to the infinite reducibility of the theories. We will briefly explain the formulation of the quantization later. We thus know how to quantize the above mentioned gravity theories. Since these gravity theories are formulated by the generalized Chern-Simons actions we naively expect that we may be able to formulate these gravity theories on the simplicial lattice manifold by simply putting $n$-form gauge fields on the $n$-simplexes of the manifold. The story is not so simple as we show the three dimensional example in the following. \subsection{Lattice Chern-Simons Gravity } \label{Action} Hereafter we concentrate on the three dimensional gravity since the formulation of the three dimensional Einstein gravity is established by the standard Chern-Simons action. Firstly it should be noted that the generalized Chern-Simons action includes the standard Chern-Simons action as a part of the full action. To make the story simpler we consider how to formulate the standard Chern-Simons action, which is the part of the generalized Chern-Simons action composed of one form only, on the simplicial manifold. We consider a three-dimensional piece-wise linear simplicial manifold which is composed of tetrahedra. In 3-dimensional Regge calculus it is known that curvature is concentrated on the links of tetrahedra\cite{Regge}. We intend to formulate a lattice gravity theory in terms of gauge variables, dreibein $e$ and spin connection $\omega$. In analogy with the lattice gauge theory where link variables surrounding a plaquette induce a gauge curvature, We have proposed the formulation that dual link variables $U(\tilde{l}) = e^{\omega(\tilde{l})}$ located at the boundary of a dual plaquette $\tilde P$ ($\tilde{l} \in \partial\tilde{P}(l)$) associated to an original link $l$ induce the curvature of the gravity theory \cite{Kawamoto-Nielsen}\cite{Dadda}. It was further pointed out that the dreibein $e^a(l)$ is located on the original link $l$. In order to avoid sign complications we consider a Euclidean version of three-dimensional local Lorentz group $SO(3)$ but exclude $SU(2)$ case. Here we explicitly construct a lattice Chern-Simons gravity by extending our previous formulation. The details of the formulation can be found in \cite{KNS}. Here we slightly modify the formulation given above in order that each tetrahedron gets independent contribution to the partition function and at the same time the orientability could be naturally accommodated. We divide the dual link, which connects the centers of neighboring tetrahedra, into two links by the center of mass of the common triangle of the neighboring tetrahedra. We may keep to use the terminology of dual plaquette and dual link even for those modified plaquettes and links. Correspondingly we put different link variables $U$ for the doubled dual links. We then assign the directions of $U$-links inward for each tetrahedron as shown in Fig.\ref{fig:action}. \begin{figure}[t] \begin{center} \begin{minipage}[c]{0.4\textwidth} \epsfxsize=\textwidth \epsfbox{action.eps} \end{minipage} \end{center} \caption{dual link variables on $\partial \tilde{P}$} \label{fig:action} \end{figure} Using these variables, we consider the following lattice version of $ISO(3)$ Chern-Simons gravity action on the simplicial manifold, \begin{equation} S_{\hbox{{\scriptsize LCS}}} = \sum_{l} \epsilon_{abc} e^{a}(l) \Bigl[ \ln \displaystyle \prod _{\partial \tilde{P}(l)} \hspace*{-6.4mm} \bigcirc \ U \Bigr]^{bc}, \label{LCS1} \end{equation} where, $\partial \tilde{P}(l)$ is a boundary of the $\tilde{P}(l)$, which is a (dual) plaquette around the link $l$, and $\displaystyle \displaystyle \prod _{\partial \tilde{P}(l)} \hspace*{-6.4mm} \bigcirc \ U $ denotes the product of $U(\tilde{l})$ along $\partial \tilde{P}(l)$. We define the ``curvature'' $F^{ab}(l)$ of the link $l$ by the following equation, \begin{equation} \Bigl[ \displaystyle \prod _{\partial \tilde{P}(l)} \hspace*{-6.4mm} \bigcirc \ U \Bigr]^{ab} \equiv \Bigl[ e^{F(l)} \Bigr]^{ab}. \end{equation} The leading term of $F$ with respect to the lattice unit is the ordinary curvature $d\omega + \omega \wedge \omega$ similar to the ordinary lattice gauge theory. Classically the Chern-Simons action impose a torsion free condition as an equation of motion. The torsion free nature is lost at the quantum level since we integrate out the dreibein and spin connection. We now introduce the following vanishing holonomy constraint which relates the dreibein and spin connection even at the quantum level: \begin{equation} \Bigl[ \displaystyle \prod _{\partial \tilde{P}(l)} \hspace*{-6.4mm} \bigcirc \ U \Bigr]^{ab} e^b = e^a. \label{constraint} \end{equation} The dreibein $e^a$ associated to a original link may be parallel transported around the boundary of the dual plaquette $\partial \tilde{P}(l)$ to the original location and yet the direction of the dreibein should not be changed. We may interpret this constraint as a gauge fixing condition of gauge diffeomorphism symmetry which we will explain later. Due to the constraint the group $SO(3)$ becomes ``effectively abelian'', i.e. the direction of the rotation associated with the curvature is parallel to that of $e^a$. This can be seen as follows: we can reduce the above constraint to the following one: \begin{equation} F^{ab} e^b = 0, \label{constraint2} \end{equation} hence $F^a \equiv \frac{1}{2} \epsilon ^{abc} F^{bc}$ is parallel to $e^a$: $e^a \propto F^a$. Here we should reconsider the constraint (\ref{constraint}). Firstly it should be noted that the $\displaystyle\Bigl[ \displaystyle \prod _{\partial \tilde{P}(l)} \hspace*{-6.4mm} \bigcirc \ U \Bigr]^{ab}$ is an element of $SO(3)$ and thus the eigenvalue equation of this element always has eigenvalue +1. Thus the number of the independent constraints in eq.(\ref{constraint}) is not three but two. Taking into account the parallel and anti-parallel nature of $e^a$ and $F^a$ in the constraint, we can rewrite the correct constraint equation \begin{equation} \frac{e^3}{|e|} \left[ \prod_{a=1}^{2} \delta \left( \frac{F^a}{|F|} + \frac{e^a}{|e|} \right) + \prod_{a=1}^{2} \delta \left( \frac{F^a}{|F|} - \frac{e^a}{|e|} \right) \right], \label{constraint3} \end{equation} where $|e|$ and $|F|$ are length of $e^a$ and $F^a$, respectively. The coefficient factor $\frac{e^3}{|e|}$ is necessary to keep the rotational invariance of the constraint relation, which can be easily checked by polar coordinate expression of the constraint relation. Now we show that discreteness of the length of the dreibein $|e|$ comes out as a natural consequence of the specific choice of the lattice gauge gravity action. We first introduce the following normalized matrix $I$, \begin{equation} I \equiv I^{a} J_{a} , \quad I^{a} \equiv \frac{F^{a}}{\sqrt{F^{a}F_{a}}}, \end{equation} here $[J_{a}]_{bc} = i \epsilon_{abc}$ is the generator of $SO(3)$. This matrix satisfies the following relation, \begin{equation} e^{i \theta I} = 1 - I^2 (1-\cos \theta) + i I \sin \theta, \end{equation} then \begin{equation} e^{i 2 \pi n I} = 1, \quad n \in \bm{Z} \label{periodicity1}. \end{equation} Using the above relation and $F^a \propto e^a$ by the constraint (\ref{constraint}), we find that our lattice Chern-Simons action $S_{\hbox{{\scriptsize LCS}}}$ has the following ambiguity: \begin{eqnarray*} S_{\hbox{{\scriptsize LCS}}} &=& \sum _{l} \epsilon _{abc} e^{a}(l) \left[ \ln e^{F(l)} \right] ^{bc} \\ &=& \sum _{l} \epsilon _{abc} e^{a}(l) \left[ \ln e^{F(l) + i 2 \pi n I} \right] ^{bc} \\ &=& \sum _{l} \bigl[ 2 e^{a}(l) F_a (l) + 4 \pi n |e(l)| \bigr] \\ &=& S_{\hbox{{\scriptsize LCS}}} + \sum_{l} 4 \pi n |e(l)|, \label{ambiguity} \end{eqnarray*} here $|e|$ is the length of $e^a$, $|e| \equiv \sqrt{e_a e^a}$. This ambiguity leads to an ambiguity in the partition function \begin{equation} Z = \int {\cal D}U {\cal D}e ~ e^{i S_{\hbox{\tiny LCS}}} = \int {\cal D}U {\cal D}e ~ e^{ i S_{\hbox{\tiny LCS}} + i \sum_{l} 4 \pi n |e|}. \end{equation} Imposing the single valuedness of $e^{i S_{\hbox{\tiny LCS}}}$, we obtain the constraint that $\sum_{l} 2 |e(l)| $ should be integer, or equivalently $|e(l)|$ should be half integer. \subsubsection{Gauge Invariance on the Lattice} \label{Gauge Invariance} The gauge transformations of the continuum Chern-Simons gravity have been given by (\ref{CS gauge transformation}) which includes the local Lorentz gauge transformation (\ref{LLGT}) and the gauge transformation of diffeomorphism (\ref{DGT}). We first note that the dreibein and the curvature defined in (\ref{CSCurvature}) transform adjointly under the local Lorentz gauge transformation \begin{equation} \begin{array}{rcl} \delta e^a_\mu &= -\epsilon^{abc}e_{\mu b}\tau_c,\\ \delta F^a_{\mu\nu} &= -\epsilon^{abc}F^b_{\mu\nu}\tau_c. \label{LLGT2} \end{array} \end{equation} We consider that the lattice version of the local Lorentz gauge parameters are sitting on the dual sites and the middle of the original links, the same point of the dreibein. For simplicity we consider here in this section that the dual link is not divided into two dual links by the center of original triangle. Then the dual link variable $U(\tilde{l}) = e^{\omega(\tilde{l})}$ transforms under the lattice local Lorentz transformation as \begin{equation} U(\tilde{l}) \rightarrow V^{-1} U(\tilde{l}) V', \label{LLLspin} \end{equation} where the gauge parameters $V$ and $V'$ are elements of $SO(3)$ and located at the end points of dual link $\tilde{l}$. Defining the matrix form of the dreibein by $E^{cb}_\mu(l)=\epsilon^{abc}e_{a\mu}(l)$, we can rewrite the lattice Chern-Simons action (\ref{LCS1}) by \begin{equation} S_{\hbox{{\scriptsize LCS}}} = \sum_{l} \hbox{Tr} ( E(l)F(l)), \label{LCSA} \end{equation} where $\displaystyle F(l)^{ab}=\Bigl[ \ln \displaystyle \prod _{\partial \tilde{P}(l)} \hspace*{-6.4mm} \bigcirc \ U \Bigr]^{ab}$. Corresponding to the continuum local Lorentz transformation, we can define the lattice version of local Lorentz transformation of $E(l)$ and $F(l)$ according to (\ref{LLLspin}) \begin{equation} \begin{array}{ccc} E(l) & \rightarrow & V^{-1}E(l) V, \\ F(l) & \rightarrow & V^{-1}F(l) V. \end{array} \label{LLLGT} \end{equation} It is obvious that the lattice Chern-Simons action (\ref{LCSA}) is invariant under the lattice local Lorentz transformation. The continuum Chern-Simons gravity action is invariant under the gauge transformation of diffeomorphism (\ref{DGT}) which transforms dreibein $e^a_\mu$ but not spin connection $\omega^a_\mu$. We can show that the lattice Chern-Simons action is invariant under the lattice version of gauge diffeomorphism by formulating the lattice version of Bianchi identity\cite{KNS}. We now point out that the constraint (\ref{constraint}) or equivalently (\ref{constraint2}) breaks the lattice gauge diffeomorphism while the lattice Chern-Simons action itself is invariant, as is shown above. The lattice dreibein is transformed but the lattice curvature is not transformed under the lattice gauge transformation of the diffeomorphism. The precise expression of the constraint (\ref{constraint3}) tells us that the dreibein $e^a$ can be rotated by using two gauge parameters of the gauge transformation of diffeomorphism to be parallel or anti-parallel to the curvature $F^a$. The length of the dreibein is discretized and thus the third gauge parameter can be exhausted. In this sense we can identify the equivalent constraint, (\ref{constraint}), (\ref{constraint2}), (\ref{constraint3}) as a gauge fixing condition of the lattice gauge transformation of diffeomorphism. \subsubsection{Calculation of Partition Function} \label{Integration} In the previous section we have found that the length of dreibein is discretized to half integer for $SO(3)$. Taking into account the discreteness of the dreibein, the total partition function leads \begin{eqnarray} Z &=& \int {\cal D}U \prod_{l} Z_{l}, \\ Z_{l} &=& \int d^3 e ~\frac{e^3}{|e|} \left[ \prod_{a=1}^{2} \delta \left( \frac{F^a}{|F|} + \frac{e^a}{|e|} \right) + \prod_{a=1}^{2} \delta \left( \frac{F^a}{|F|} - \frac{e^a}{|e|} \right) \right] \nonumber \\ && \hspace*{1cm} \times \frac{1}{2}\sum_{J=0}^{\infty} \delta \left( |e| - \frac{J}{2} \right) e^{2 i e^a F^a}, \end{eqnarray} where $Z_{l}$ is the partition function associated with a link $l$. {\bf 4.2.2.1 $e$ integration}\\ Due to the rotational invariance of the constraints, we can take $e^3$ as the third direction of local Lorentz frame without loss of generality. We can then evaluate $e^a$ integral of $Z_{l}$ immediately thanks to the delta functions \begin{eqnarray*} Z_{l} &=& \int d^3 e ~ |e|^2 \frac{e^3}{|e|} \left[ \prod_{a=1}^{2} \delta \left( e^a + |e| \frac{F^a}{|F|} \right) + \prod_{a=1}^{2} \delta \left( e^a - |e| \frac{F^a}{|F|}\right) \right] \frac{1}{2}\sum_{J} \delta \left( |e| - \frac{J}{2} \right) e^{2 i e^a F^a} \\ &=& \frac{1}{2}\sum_{J} \left( \frac{J}{2} \right)^2 \left( e^{2i\frac{J}{2} |F|} + e^{-2i\frac{J}{2} |F|} \right) \\ &=& \sum_{J} \frac{1}{4} J^2 \cos(J|F|). \end{eqnarray*} Using the following formula for the character $\chi_J$ of the spin-$J$ representation of $SO(3)$, \begin{equation} \chi _{J} (e^{i \theta^a J_a}) = \chi _{J} (|\theta|) = \frac{\sin\left( (2J+1) \frac{|\theta|}{2} \right)} {\sin \left(\frac{|\theta|}{2} \right)}, \end{equation} where $|\theta|$ is the length of $\theta^a$, we find \begin{eqnarray} \chi _{J}(|F|) - \chi _{J-1}(|F|) &=& 2 \cos(J|F|). \end{eqnarray} Hence we can naively calculate the link partition function, \begin{eqnarray*} Z_{l} &=& \sum_{J=1}^{\infty} \frac{1}{8} J^2 (\chi _{J} - \chi _{J-1}) \\ &=& -\frac{1}{8} \sum_{J=0}^{\infty} (2J+1) \chi _{J}. \end{eqnarray*} This calculation is not precise, because the summation is not convergent. We need to show that there is a regularization procedure which leads to a validity of the above calculation after the regularization. It is possible to give the similar formulation by using Heat Kernel regularization. Then the action leads \begin{eqnarray} Z_{l} &=& - \frac{1}{8} \sum_{J} (2J+1)\chi _J e^{-J(J+1)t}. \end{eqnarray} This regularization factor $e^{-J(J+1)t}$ breaks the Alexander move invariance of the partition function but it will be recovered at the end of the calculation when we take the limit $t \rightarrow 0$. {\bf 4.2.2.2 $U$ integration}\\ After $e^a$ integration and dividing the unimportant constant factor $\prod_{l} (-1/8)$, the partition function leads \begin{equation} Z = \int {\cal D}U \prod_{l} \sum_{J=0}^{\infty} \left(2J+1\right) \chi_{J}\Bigl(\displaystyle \prod _{\partial \tilde{P}(l)} \hspace*{-6.4mm} \bigcirc \ U \Bigr) ~ e^{-J(J+1)t}, \end{equation} where we take $t\rightarrow 0$ limit in the end of calculation. We now carry out $DU$ integration of this partition function. Thanks to the character of the partition function, $DU$ integration is straightforward. We show that the Ponzano-Regge partition function will be reproduced after $DU$ integration with 6-$j$ symbols together with correct coefficients and sign factors. Before getting into the details we figure out how 6-$j$ symbols appear. The character in the partition function is a product of $D$-function around the boundary of dual plaquette associated to a original link. Each tetrahedron has six original links and there are two dual links which is a part of a product on the boundary of the dual plaquette associated to each original link. In other words three dual links associated to a $DU$ integration thrust into each triangle from the center of the tetrahedron. Therefore twelve dual links are associated to a tetrahedron. Each $DU$ integration of the product of three $D$-function reproduces two 3-$j$ symbols, thus we get eight 3-$j$ symbols for each tetrahedron. Four out of eight 3-$j$ symbols lead to a 6-$j$ symbol and the rest of four 3-$j$ symbols lead to give a trivial factor together with the 3-$j$ symbols from the neighboring tetrahedra. We first note that the character appearing in the partition function is a product of $D$-functions \begin{equation} \chi_J(|F|) = \chi_J \Bigl(\displaystyle \prod _{\partial \tilde{P}(l)} \hspace*{-6.4mm} \bigcirc \ U \Bigr) = D^J_{m_1m_2}(U_1)D^J_{m_2m_3}(U_2)\cdots D^J_{m_km_1}(U_k), \end{equation} where $U_i$ is a dual link variables on the boundary of dual plaquette $\tilde P(l)$ associated to a link $l$ and $m_i$ is the third component of angular momentum $J$ which is assigned to the link $l$. As we have already pointed out that the direction of $U_i$ for each link is defined inward for each tetrahedron. On the other hand the direction of the loop composed of the product of dual links associated to the link $l$ can be chosen arbitrarily. Therefore some of $U_i$ in the above $D$-functions are $U_i^\dagger$. If the original link $l$ is a link of a particular tetrahedron, two $D$-functions out of the above product are located inside the tetrahedron. \begin{figure} \begin{center} \begin{minipage}[b]{0.4\textwidth} \epsfxsize=\textwidth \epsfbox{U-int.eps} \end{minipage} \hspace*{5mm} \begin{minipage}[b]{0.4\textwidth} \epsfxsize=\textwidth \epsfbox{U-int1.eps} \end{minipage} \end{center} \caption{dual links related to neighboring tetrahedra and the orientability} \label{fig:U-int} \end{figure} We now choose a particular situation which is shown in Fig.\ref{fig:U-int}. The twelve $D$-functions associated to this particular tetrahedron are \begin{eqnarray*} && I_{U_1U_2U_3U_4} = \int \prod_{i=1}^{4} DU_i ~ D^{J_1}_{i_1m_1}(U_1) D^{J_1}_{m_1k_3}(U_3^{\dagger}) \cdot D^{J_2}_{j_2m_2}(U_2) D^{J_2}_{m_2i_2}(U_1^{\dagger}) \\ && \quad\quad\quad\quad \times D^{J_3}_{l_1m_3}(U_4) D^{J_3}_{m_3i_3}(U_1^{\dagger}) \cdot D^{J_4}_{l_2m_4}(U_4) D^{J_4}_{m_4j_1}(U_2^{\dagger}) \\ && \quad\quad\quad\quad \times D^{J_5}_{l_3m_5}(U_4) D^{J_5}_{m_5k_1}(U_3^{\dagger}) \cdot D^{J_6}_{k_2m_6}(U_3) D^{J_6}_{m_6j_3}(U_2^{\dagger}). \end{eqnarray*} We pick up the $D$-functions associated to $DU_1$ integration \begin{equation} I_{U_1}= (-)^{i_2-m_2+i_3-m_3} \int DU_1 D^{J_1}_{i_1m_1}(U_1)D^{J_2}_{-i_2-m_2}(U_1) D^{J_3}_{-i_3-m_3}(U_1), \end{equation} where we have used the following formula to rewrite only with $U_1$ variable: \begin{equation} D^I_{mn} (U^{\dagger}) = D^{I *}_{nm} (U) = (-)^{n-m} D^I_{-n-m} (U) \label{D conjugate}. \end{equation} We can now use the formula relating the integration of three $D$-functions and two 3-$j$ symbols given in (\ref{eqn:3d3j}) and obtain \begin{equation} I_{U_1}= (-)^{i_2-m_2+i_3-m_3} \threej{J_1}{J_2}{J_3}{m_1}{-m_2}{-m_3} \threej{J_1}{J_2}{J_3}{i_1}{-i_2}{-i_3}. \end{equation} After carrying out $DU_2DU_3DU_4$ integration, we obtain \begin{eqnarray*} I_{U_1U_2U_3U_4} &=& (-)^{i_2-m_2+i_3-m_3} \threej{J_1}{J_2}{J_3}{m_1}{-m_2}{-m_3} \threej{J_1}{J_2}{J_3}{i_1}{-i_2}{-i_3} \\ &\times& (-)^{j_1-m_4+j_3-m_6} \threej{J_4}{J_2}{J_6}{-m_4}{m_2}{-m_6} \threej{J_4}{J_2}{J_6}{-j_1}{j_2}{-j_3} \\ &\times& (-)^{k_3-m_1+k_1-m_5} \threej{J_1}{J_5}{J_6}{-m_1}{-m_5}{m_6} \threej{J_1}{J_5}{J_6}{-k_3}{-k_1}{k_2} \\ &\times& \hspace{3cm} \threej{J_4}{J_5}{J_3}{m_4}{m_5}{m_3} \threej{J_4}{J_5}{J_3}{l_2}{l_3}{l_1}. \end{eqnarray*} We now use the formula given in (\ref{eqn:6j3j}) which relates 6-$j$ symbol and four 3-$j$ symbols which carry $m_i$ suffices associated to the center of the tetrahedron. We then find 6-$j$ symbols after $DU_1DU_2DU_3DU_4$ integration \begin{eqnarray} I_{U_1U_2U_3U_4} &=& (-)^{\sum_{i=1}^{6} J_i} \sixj{J_1}{J_2}{J_3}{J_4}{J_5}{J_6} \nonumber \\ &\times& (-)^{i_2+i_3} \threej{J_1}{J_2}{J_3}{i_1}{-i_2}{-i_3} (-)^{j_3+j_1} \threej{J_4}{J_2}{J_6}{-j_1}{j_2}{-j_3} \nonumber \\ &\times& (-)^{k_3+k_1} \threej{J_1}{J_5}{J_6}{-k_3}{-k_1}{k_2} \threej{J_4}{J_5}{J_3}{l_2}{l_3}{l_1}. \label{6jwith3j} \end{eqnarray} Here we are considering $SO(3)$ case then the third component of the angular momentum $m_i$ is integer and thus we can use the relation $(-)^{m_i}=(-)^{-m_i}$. We now look at the rest of the 3-$j$ symbols in eq.(\ref{6jwith3j}) which carry the suffices $i,j,k,l$. As we can see from Fig.\ref{fig:U-int} that $DU_1$ integration reproduces two 3-$j$ symbols and one of them associated to the suffices $m_k$ is absorbed to reproduce the 6-$j$ symbol and the other 3-$j$ symbol carrying the suffices $i_k$ could be combined with another 3-$j$ symbol obtained from $DU'_1$ integrations of the neighboring tetrahedron. Those 3-$j$ symbols are associated to the boundary triangle of the two neighboring tetrahedra carrying suffix $i_k$. In this particular case of Fig.\ref{fig:U-int} we obtain the following two 3-$j$ symbols \begin{eqnarray*} I^b_{i} = \sum_{i_1i_2i_3}(-)^{i_1+i_2+i_3} \threej{J_1}{J_2}{J_3}{i_1}{-i_2}{-i_3} \threej{J_1}{J_2}{J_3}{i_1}{-i_2}{-i_3}. \end{eqnarray*} Since the three angular momentum vectors $J_1,J_2,J_3$ construct the boundary triangle, the third components satisfy the relation $i_1-i_2-i_3=0$. Using the following formula: \begin{equation} \sum_{m_1m_2m_3} \threej{J_1}{J_2}{J_3}{m_1}{m_2}{m_3} \threej{J_1}{J_2}{J_3}{m_1}{m_2}{m_3} = 1, \end{equation} and noting $(-)^{i_1+i_2+i_3}=(-)^{i_1-i_2-i_3}=1$ for $SO(3)$ case, these two 3-$j$ symbols lead to a trivial factor. Thus all the terms together lead to the Ponzano-Regge partition function. \subsubsection{The Continuum Limit of the Lattice Chern-Simons Gravity } We have explicitly shown that the partition function of the $ISO(3)$ lattice Chern-Simons action exactly coincides with the Ponzano-Regge model after the integration of the dreibein and the dual link variables. The discreteness of the length of the dreibein is the natural consequence of the logarithm form in the lattice Chern-Simons action. On the simplicial lattice manifold constructed from tetrahedra, the drei beins are located on the original links while the lattice version of the spin connection, the dual link variables are located on the dual links. Since the Ponzano-Regge model is invariant under the 2-3 and 1-4 Alexander moves, the partition function is invariant under how the three dimensional space is divided into small pieces by tetrahedra. It is natural to expect that the partition function is invariant in the continuum limit and thus the lattice Chern-Simons action leads to the continuum Chern-Simons action. In the $ISO(3)$ lattice Chern-Simons action there are 6 gauge parameters. Two gauge parameters of the lattice gauge diffeomorphism can be used to rotate the dreibein $e^a$ to be parallel or anti-parallel to the curvature $F^a$ and the remaining one gauge parameter of the lattice gauge diffeomorphism can be exhausted to make the length of the dreibein discrete. There remain three gauge parameters of the lattice local Lorentz gauge symmetry, which are expected to convert into the three vector parameters of general coordinate diffeomorphism symmetry. There are two reasons to expect this scenario. Firstly the lattice action coincides with the Ponzano-Regge model which is Alexander move invariant and is thus expected to be metric independent. In fact the lattice Chern-Simons action in the continuum limit is metric independent since it is composed of one form. Secondly the general coordinate transformation of diffeomorphism and the local Lorentz transformation are on shell equivalent in the continuum $ISO(3)$ Chern-Simons gravity\cite{Witten1}. It is interesting to recognize that the drei bein and spin connection are located on the original and dual links, respectively. This geometrical dual nature will be the reflection of the algebraic dual nature of the $ISO(3)$ Chern-Simons gravity where the abelian momentum generators $P_a$ and the angular momentum generator $J_a$ are algebraically dual to each other\cite{Ashtekar-Romano}. \setcounter{equation}{0} \section{Quantization of Generalized Gauge Theory} Since the generalized gauge theory has nontrivial algebraic structures, we need to understand the origin of these algebras. For example the quaternion structure plays a very crucial role to treat the generalized fields and parameters in a uniform way. We expected that the quaternion structure will be essentially related with the algebraic structure of the quantization of the theories. This expectation has turned out to be true. We found several other interesting features of the generalized gauge theory in the quantization procedure. Here in this section we briefly summarize how to quantize the generalized Chern-Simons action\cite{KSTU}. In particular we concentrate on the quantization of even dimensional generalized Chern-Simons actions. To be concrete for the notations we focus on the four dimensional case here. The details of the quantization of two dimensional model is given in \cite{KSTU}. The extension to odd dimensions goes quite parallel to the even dimensional case. We consider to quantize the even dimensional generalized Chern-Simons action \begin{equation} \displaystyle{S^e_{GCS} = \int_M \mbox{Tr}_{\bf k} \left( \frac{1}{2}{\cal A}Q{\cal A}+\frac{1}{3}{\cal A}^3 \right)}, \label{eqn:egcs2} \end{equation} where the explicit component wise notations for the classical gauge fields and gauge parameters are given in (\ref{eqn:def-gf}) and (\ref{eqn:def-gp}). An important characteristic of the generalized Chern-Simons action is the infinite reducibility of the action which can be stated as follows. We extend the generalized gauge field and parameter to accommodate the ghost and the ghost for ghost $\cdots$ as \begin{eqnarray} {\cal{V}}_{2n} & = & {\mbox{\bf j}} \left(u_{2n} + U_{2n}\right) + {\mbox{\bf k}} \left( v_{2n} + b_{2n} + V_{2n}\right) \ \in \Lambda_{-}, \label{eqn:Ve4}\\ {\cal{V}}_{2n+1} & = & {\mbox{\bf 1}} \left( v_{2n+1} + b_{2n+1} + V_{2n+1}\right) - {\mbox{\bf i}}\, \left(u_{2n+1} + U_{2n+1} \right) \ \in \Lambda_{+}, \label{eqn:Vo4}\\ & & \hspace{9cm} n = 0, 1, 2, \cdots, \nonumber \end{eqnarray} where ${\cal V}_0 \equiv {\cal A}$ and ${\cal V}_1 \equiv {\cal V}$ and the parameters with $n\ (>1)$ correspond the $n$-th reducibility parameters. The infinite reducibility of the generalized Chern-Simons actions can be seen by the transformation of ${\cal{V}}_n$ \begin{equation} \delta {\cal{V}}_n = (-)^n [ \ Q + {\cal{A}} \ , \ {\cal{V}}_{n+1} \ ]_{(-)^{n+1}}, \qquad n = 0,1,2,\cdots \label{eqn:ir} \end{equation} which satisfies the on-shell relation \begin{eqnarray} \delta(\delta{\cal{V}}_n) & = & \delta {\cal{V}}_n\Bigr\vert_{{\cal V}_{n+1}\rightarrow {\cal V}_{n+1}+\delta{\cal V}_{n+1}} - \delta {\cal{V}}_n \nonumber \\ & = & 0, \label{eqn:irr} \end{eqnarray} where we have used the equation of motion of generalized Chern-Simons action. In the above equations, $[ \ , \ ]_{(-)^n}$ is a commutator for odd $n$ and an anticommutator for even $n$. Since the transformation (\ref{eqn:ir}) for $n=0$ represents the gauge transformation, eq.(\ref{eqn:irr}) implies that the gauge transformation is infinitely reducible. In order to quantize this kind of system we need to introduce infinite series of ghosts as generalized fields \begin{eqnarray} & & C_n, \ \ C_{n\mu}, \ \ \frac{1}{2!}C_{n\mu\nu}, \ \ \frac{1}{3!}C_{n\mu\nu\rho}, \ \ \frac{1}{4!}C_{n\mu\nu\rho\sigma}, \label{eqn:4d-cnc} \\ & & \hspace{8cm} n=0, \pm 1, \pm 2, \cdots, \pm\infty, \nonumber \end{eqnarray} where the index $n$ indicates the ghost number of the fields and the fields with even (odd) ghost number are bosonic (fermionic). The fields with ghost number 0 are the classical fields $$ C_0=\phi, \ \ C_{0\mu}=\omega_\mu, \ \ C_{0\mu\nu}=B_{\mu\nu}, \ \ C_{0\mu\nu\rho}=\Omega_{\mu\nu\rho}, \ \ C_{0\mu\nu\rho\sigma}=H_{\mu\nu\rho\sigma}. $$ Then we redefine a generalized gauge field \begin{eqnarray} \widetilde{{\cal A}} &=& {\mbox{\bf 1}}\psi + {\mbox{\bf i}} \hat{\psi} + {\mbox{\bf j}} A + {\mbox{\bf k}} \hat{A} \ \in \Lambda_-, \label{eqn:4d-ggf} \\ &&\psi = \sum_{n = -\infty}^{\infty} \Big(C_{2n+1 \mu} dx^{\mu}+ \frac{1}{3!}C_{2n+1\mu\nu\rho} dx^\mu\wedge dx^\nu\wedge dx^\rho \Big), \nonumber \\ &&\hat{\psi} = \sum_{n = -\infty}^{\infty} \Big( C_{2n+1} + \frac{1}{2!} C_{2n+1 \mu \nu} dx^{\mu}\wedge dx^{\nu} + \frac{1}{4!}C_{2n+1 \mu\nu\rho\sigma} dx^\mu\wedge dx^\nu\wedge dx^\rho\wedge dx^\sigma\Big), \nonumber \\ &&A = \sum_{n = -\infty}^{\infty} \Big(C_{2n \mu} dx^{\mu}+ \frac{1}{3!}C_{2n\mu\nu\rho} dx^\mu\wedge dx^\nu\wedge dx^\rho \Big), \nonumber \\ &&\hat{A} = \sum_{n = -\infty}^{\infty} \Big( C_{2n} + \frac{1}{2!} C_{2n \mu \nu} dx^{\mu}\wedge dx^{\nu} + \frac{1}{4!}C_{2n \mu\nu\rho\sigma} dx^\mu\wedge dx^\nu\wedge dx^\rho\wedge dx^\sigma\Big), \nonumber \end{eqnarray} where we have explicitly shown the differential form dependence. Using the above definitions of generalized gauge fields, we define the following extended generalized action which has the same form as the original one \begin{equation} \displaystyle{S_{min}= \int \ {\mbox{Tr}}^0_{\mbox{\bf {k}}} \left( \frac {1}{2}\widetilde{{\cal A}}Q\widetilde{{\cal A}} + \frac {1}{3} \widetilde{{\cal A}}^3 \right)}, \label{eqn:mina2} \end{equation} where the upper index 0 on ${\mbox{Tr}}$ indicates to pick up only the part with ghost number 0. We can then show that this action satisfies the master equation of antibracket formalism \`{a} la Batalin and Vilkovisky\cite{BV}. In the construction of Batalin and Vilkovisky, ghosts, ghosts for ghosts$\cdots$ and the corresponding antifields are introduced according to the reducibility of the theory. We denote a minimal set of fields by $\Phi^A$ which include classical fields and ghost fields, and the corresponding antifields by $\Phi_A^{\ast}$. If a field has ghost number $n$, its antifield has ghost number $-n-1$. Then a minimal action is obtained by solving the master equation \begin{eqnarray} (S_{min}(\Phi , \Phi^*),S_{min}(\Phi , \Phi^*)) &=& 0, \label{eqn:me} \\ (X,Y)&=&{\partial_r X \over \partial \Phi^A} {\partial_l Y \over \partial \Phi_A^*} -{\partial_r X \over \partial \Phi_A^*} {\partial_l Y \over \partial \Phi^A}, \label{eqn:ab} \end{eqnarray} with the boundary conditions \begin{eqnarray} S_{min} \Bigr\vert_{\Phi^{\ast}_A = 0} & = & S_0, \label{eqn:bbc} \\ \frac{\partial S_{min}}{\partial \Phi^{\ast}_{a_n}} \Bigr\vert_{\Phi^{\ast}_A = 0} & = & Z^{a_n}_{a_{n+1}} \Phi^{a_{n+1}}, \label{eqn:bbd} \hspace{2cm} n = 0,1,2,\cdots, \end{eqnarray} where $S_0$ is the classical action and $Z^{a_n}_{a_{n+1}} \Phi^{a_{n+1}}$ represents the $n$-th reducibility transformation where the reducibility parameters are replaced by the corresponding ghost fields. The BRST transformation and the nilpotency of the transformation can be shown by using generalized fields \begin{eqnarray*} \delta_{\lambda} \widetilde{{\cal A}} &\equiv& s \widetilde{{\cal A}} \lambda = -\widetilde{{\cal F}} \ {\mbox{\bf i}} \lambda, \\ s^2\widetilde{{\cal A}}\lambda_2\lambda_1 &\equiv& \delta_{\lambda_2} \delta_{\lambda_1} \widetilde{{\cal A}} = - [ \ Q + \widetilde{{\cal A}} \ , \ \widetilde{{\cal F}} \ ] \lambda_2 \lambda_1 = 0, \end{eqnarray*} where the last relation holds due to the generalized Bianchi identity. We can then show that $S_{min}$ satisfies the master equation $$ \delta_{\lambda} S_{min} = (S_{min},S_{min})_{\lambda,{\bf k}} = ( S_{min} , S_{min} ) \cdot \lambda = 0, $$ where $(~,~)$ is the original antibracket defined by (\ref{eqn:me}) with the following identifications of antifields: \begin{eqnarray} \frac{1}{4!}\epsilon^{\mu\nu\rho\sigma}C_{-2n+1 \mu\nu\rho\sigma} &=&C^*_{2(n-1)}, \hspace{1.5cm} \frac{1}{4!}\epsilon^{\mu\nu\rho\sigma}C_{-2n \mu\nu\rho\sigma} =-C^*_{2n-1}, \nonumber \\ \frac{1}{3!}\epsilon^{\mu\nu\rho\sigma}C_{-2n+1 \nu\rho\sigma} &=&C^{\mu*}_{2(n-1)}, \hspace{1.5cm} \frac{1}{3!}\epsilon^{\mu\nu\rho\sigma}C_{-2n \nu\rho\sigma} =C^{\mu*}_{2n-1}, \nonumber \\ \frac{1}{2!}\epsilon^{\mu\nu\rho\sigma}C_{-2n+1 \rho\sigma} &=&-C^{\mu\nu*}_{2(n-1)}, \hspace{1.5cm} \frac{1}{2!}\epsilon^{\mu\nu\rho\sigma}C_{-2n \rho\sigma} =-C^{\mu\nu*}_{2n-1}, \label{eqn:4d-idantif} \\ \epsilon^{\mu\nu\rho\sigma}C_{-2n+1 \sigma} &=&C^{\mu\nu\rho*}_{2(n-1)}, \hspace{1.5cm} \epsilon^{\mu\nu\rho\sigma}C_{-2n \sigma} =C^{\mu\nu\rho*}_{2n-1}, \nonumber \\ \epsilon^{\mu\nu\rho\sigma}C_{-2n+1} &=&C^{\mu\nu\rho\sigma*}_{2(n-1)}, \hspace{1.5cm} \epsilon^{\mu\nu\rho\sigma}C_{-2n } =-C^{\mu\nu\rho\sigma*}_{2n-1}. \nonumber \end{eqnarray} In order to fix gauge completely we need to introduce the non-minimal action with the proper choice of gauge fermion. We can then eliminate the antifields via the chosen gauge fermion. We just mention here that all these procedures can be completed properly\cite{KSTU}. We can thus complete the quantization of the generalized Chern-Simons action in even dimensions. \setcounter{equation}{0} \section{Generalized Yang-Mills Theory} Interesting applications of the generalized gauge theory would be to find the realistic correspondence with the known formulations. Needless to say that the Yang-Mills action has been playing a crucial role to describe the three out of four fundamental forces of our nature and may play again an important role in describing unified theory of all the four fundamental forces including gravity. Here in this section we concretely show that the generalized Yang-Mills action is fundamentally related to the super symmetry and interpretation of matter fermions\cite{Kawamoto-Tsukioka} and non-commutative geometry formulation of gauge theory\cite{KTU}. \subsection{Generalized Topological Yang-Mills Theory} In this subsection we analyze the two dimensional version of the generalized topological Yang-Mills action and show that the instanton gauge fixing of the generalized topological Yang-Mills action leads to a $N=2$ super Yang-Mills theory together with the twisting procedure. In four dimensions Witten showed that a topological gauge theory having $N=2$ super symmetry leads to the $N=2$ super Yang-Mills theory with the twisting procedure\cite{Witten2}. Later it has been pointed out that the $N=2$ super Yang-Mills theory with the twisting procedure can be derived from the topological Yang-Mills action with instanton gauge fixing\cite{Baulieu-Singer}. This subject has been intensively investigated \cite{top-twist} in particular the twisting mechanism has been cleared up in two dimensions with the help of super conformal field theory\cite{Eguchi-Yang}. Our formulation of this section\cite{Kawamoto-Tsukioka} is the two dimensional realization of the known four dimensional scenario and can be extended to arbitrary dimensions. To make the formulation concrete and simpler we specify to the two dimensional case. As we have already mentioned that the action we consider satisfies the following well known relation: \begin{equation} \int_M{\mbox{Str}}_{\bf\mbox{1}}{\cal F}^2_0 =\int_M{\mbox{Str}}_{\bf\mbox{1}}\Bigg( Q\Big({\cal{A}}_0Q{\cal{A}}_0 +\frac{2}{3}{\cal A}_0^3\Big)\Bigg), \label{eq:gym2} \end{equation} where ${\cal A}_0$ and ${\cal F}_0$ are the two dimensional counter part of the generalized classical gauge field and curvature. More explicitly they are given by \begin{eqnarray} {\cal{A}}_0 &=& {\bf\mbox{j}}\omega + {\bf\mbox{k}}\Big( \phi +B \Big) \qquad\in\Lambda_-, \\ {\cal{F}}_{0} &=& Q{\cal{A}}_0 + {\cal{A}}_0^2 \\ &=& -{\bf\mbox{1}}\Big( d\omega + \omega^2 + \{ \phi, B \} +\phi^2 \Big) +{\bf\mbox{i}}\Big(d\phi+[\omega, \phi] \Big) \qquad\in\Lambda_+. \end{eqnarray} Due to the topological nature of the action, i.e., the action vanishes identically if the two dimensional manifold $M$ does not have boundary, the action has so called shift symmetry. In other words the action is invariant under the gauge transformation of an arbitrary function ${\cal{E}}_0$. Thus the gauge transformation of the generalized topological Yang-Mills action has the following form: \begin{equation} \delta {\cal A}_{0} = [Q+{\cal{A}}_0, {\cal{V}}_0] +{\cal{E}}_0, \label{eqn:ty-gt} \end{equation} where ${\cal{V}}_0$ is the generalized gauge parameter \begin{equation} {\cal{V}}_0 = {\bf\mbox{1}}\Big(v +b \Big) +{\bf\mbox{i}}u \qquad\in\Lambda_+, \end{equation} while ${\cal{E}}_0$ is a new gauge parameter of the shift symmetry and is given by \begin{equation} {\cal{E}}_0 = {\bf\mbox{j}}\xi^{(1)} + {\bf\mbox{k}}\Big(\xi^{(0)} +\xi^{(2)}\Big) \qquad\in\Lambda_-, \end{equation} where $\xi^{(0)},\xi^{(1)}$ and $\xi^{(2)}$ are zero-, one- and two form bosonic gauge parameters, respectively. At this stage the readers may wonder why we should keep the original gauge transformation in the gauge transformation of ${\cal A}_{0}$ in (\ref{eqn:ty-gt}) since the gauge parameter of the shift symmetry ${\cal{E}}_0$ is an arbitrary function and could absorb the change of the original gauge transformation. It will turn out later that both of the shift symmetry and the original symmetry are essential to induce $N=2$ super symmetry with matter fermions via twisting procedure. The generalized topological action has the following obvious reducibility: \begin{equation} \begin{array}{rcl} {\cal{V}}_0 &=& {\cal{V}}_1, \\ {\cal{E}}_0 &=& -[Q+{\cal{A}}_0, {\cal{V}}_1]. \label{eqn:ty-reducibility} \end{array} \label{eq:reducibility} \end{equation} Thus this system is a first stage reducible system in the terminology of Batalin and Vilkovisky. Correspondingly we need to introduce ghost fields $C^{(n)}$ and $\widetilde{C}^{(n)}$ with respect to the gauge parameters ${\cal{V}}_0$ and ${\cal{E}}_0$, and ghost for ghost field $\eta^{(n)}$ with respect to ${\cal{V}}_1$ , respectively. Here the suffix $(n)$ with $n=0,1,2$ denotes the form degree. We then redefine the generalized gauge field by introducing the ghost field $C^{(n)}$ \begin{equation} {\cal{A}} = {\bf\mbox{1}}C^{(1)} +{\bf\mbox{i}}\Big(C^{(0)} +C^{(2)}\Big) +{\bf\mbox{j}}\omega +{\bf\mbox{k}}\Big(\phi + B\Big) \qquad\in\Lambda_-. \label{eqn:gfa} \end{equation} We need to introduce another generalized field to accommodate the ghost of the shift symmetry $\widetilde{C}^{(n)}$ and ghost for ghost $\eta^{(n)}$ \begin{equation} {\cal{C}} = {\bf\mbox{1}}\Big(\eta^{(0)} + \eta^{(2)} \Big) +{\bf\mbox{i}}\eta^{(1)} +{\bf\mbox{j}}\Big(\widetilde{C}^{(0)} + \widetilde{C}^{(2)}\Big) +{\bf\mbox{k}}\widetilde{C}^{(1)} \qquad\in\Lambda_+. \label{eqn:gfc} \end{equation} Here ${\cal{C}}$ belongs $\Lambda_+$ and could be identified as a part of generalized curvature later. Furthermore we extend the differential operator $Q$ by introducing the BRST operator $s$ as a fermionic zero-form \begin{equation} {\cal{Q}} \equiv Q+Q_B = {\bf\mbox{j}}d+{\bf\mbox{i}}s, \qquad\in\Lambda_-. \end{equation} It should be noted that $s$ commutes with $d$, $i.e.$ $[d, s]=0$ and $s^2=0$. This operator satisfies the nilpotency property because of quaternionic structures \begin{equation} {\cal{Q}}^2 = 0. \label{eq:np} \end{equation} The following graded Leibnitz rule acting on generalized gauge fields ${\cal{A}}, {\cal{B}}\in\Lambda_\pm$ can be derived: \begin{equation} {\cal{Q}}({\cal{A}}{\cal{B}}) =({\cal{Q}}{\cal{A}}){\cal{B}} +(-)^{|{\cal{A}}|}{\cal{A}}({\cal{Q}}{\cal{B}}), \label{eqn:lr} \end{equation} where $|{\cal{A}}|=0$ for ${\cal{A}}\in\Lambda_+$ and $|{\cal{A}}|=1$ for ${\cal{A}}\in\Lambda_-$. We can now define the generalized curvature by using the redefined gauge field \begin{eqnarray} {\cal{F}} &=& {\cal{Q}}{\cal{A}}+{\cal{A}}^2 \nonumber \\ &=& {\cal{F}}_0 + {\cal{C}}, \label{eqn:gecu} \end{eqnarray} where the second relation is imposed to relate the BRST transformation of the components fields. The nilpotency of the BRST transformation is assured by the Bianchi identity of the generalized field \begin{equation} {\cal{Q}}{\cal{F}}+[{\cal{A}}, {\cal{F}}]=0. \label{eqn:geBi} \end{equation} The component wise expressions of BRST transformation can be read from (\ref{eqn:gecu}) and (\ref{eqn:geBi}). \subsubsection{Instanton Gauge Fixing of Topological Yang-Mills Model} We next introduce a particular model to carry out explicit analyses. In particular we choose the Clifford algebra as a graded Lie algebra, which closes under the multiplication and is simplest nontrivial example. More specifically we take the following two dimensional antihermitian Euclidean Clifford algebra: \begin{equation} \begin{array}{rcl} T^a &:& 1, \quad \gamma_5, \\ \Sigma^\alpha &:& \gamma^a, \end{array} \end{equation} where $\gamma^a=(i\sigma^1, i\sigma^2)$, which satisfy $\{ \gamma^a, \gamma^b\}=-2\delta^{ab}$ and $\displaystyle\gamma_5=\frac{1}{2}\epsilon_{ab}\gamma^a\gamma^b =-i\sigma^3$ with $\epsilon_{12}=1$. A grading generator can be identified by $\gamma_5$ and then we define the super trace $$ {\mbox{Str}}(\cdots)=\hbox{Tr}(\gamma_5\cdots). $$ The two dimensional generalized topological Yang-Mills action leads \begin{eqnarray} S_0&=&\frac{1}{2}\int_M{\mbox{Str}}_{\bf\mbox{1}}{\cal F}_0^2, \nonumber \\ &=&\int_M d^2x\Big(\epsilon^{\mu\nu}F_{\mu\nu}|\phi|^2 +\epsilon^{\mu\nu}\epsilon^{ab}(D_\mu\phi)_a (D_\nu\phi)_b \Big) \nonumber \\ &=&\int_M d^2x\epsilon^{\mu\nu}\partial_\mu \Big(2\omega_\nu|\phi|^2+ \epsilon^{ab}\phi_a\partial_\nu\phi_b\Big), \label{eqn:2ty} \end{eqnarray} where the scalar part of the one-form field $\omega_{\mu s}$ and the two form field $B_{a\mu\nu}$ do not appear in the action. It can be seen from the above relation that the square of the generalized curvature is related to the one dimensional generalized Chern-Simons action with the particular choice of the graded Lie algebra. We can now find out a two dimensional instanton relation of our generalized topological Yang-Mills action by imposing self- (anti-self-) dual condition \begin{equation} *{\cal F}_0=\pm{\cal F}_0. \label{eqn:ic} \end{equation} In the present model we take the following duality relation for the gauge algebra and the quaternions \begin{eqnarray} *1 = -\gamma_5, ~~~ {*}\gamma^a = -{\epsilon^a}_b\gamma^b, ~~~ {*}\gamma_5 = -1, \nonumber\\ {*}{\bf\mbox 1}={\bf\mbox 1}, ~~~~~~~~~~~~~~~ {*}{\bf\mbox i}=-{\bf\mbox i}. \end{eqnarray} We can then find the following minimal condition of the action leading to instanton relations: \begin{eqnarray} & &\pm\frac{1}{2}\int\hbox{Str}_{\bf\mbox{1}}{\cal F}_0\wedge{\cal F}_0 +\frac{1}{2}\int\hbox{Str}_{\bf\mbox{1}}{\cal F}_0\wedge*{\cal F}_0 \nonumber \\ &=& \frac{1}{4}\int\hbox{Str}_{\bf\mbox{1}} \Big({\cal{F}}_0\pm*{\cal F}_0\Big)\wedge *\Big({\cal F}_0\pm*{\cal F}_0\Big) \nonumber \\ &=&\int d^2x\Bigg(\Big(\frac{1}{2} \epsilon^{\mu\nu}F_{\mu\nu}\pm|\phi|^2\Big) \Big(\frac{1}{2} \epsilon^{\rho\sigma}F_{\rho\sigma}\pm|\phi|^2\Big) \nonumber \\ & &\qquad\quad +\frac{1}{2}\Big((D_\mu\phi)_a \pm{\epsilon_\mu}^\nu {\epsilon_a}^b (D_\nu\phi)_b\Big) \Big((D^\mu\phi)^a \pm{\epsilon^\mu}_\rho {\epsilon^a}_c (D^\rho\phi)^c\Big)\Bigg). \label{eqn:tpi} \end{eqnarray} Then the instanton condition is the absolute minima of the generalized Yang-Mills action \begin{eqnarray} \frac{1}{2}\epsilon^{\mu\nu}F_{\mu\nu}-|\phi|^2 &=& 0, \label{eq:sdgf1} \\ (D_\mu\phi)_a^{(-)}&\equiv&\displaystyle{\frac{1}{2} \Big((D_\mu\phi)_a-{\epsilon_\mu}^\nu {\epsilon_a}^b(D_\nu\phi)_b \Big)} = 0. \label{eqn:sdgf2} \end{eqnarray} We now derive the gauge fixed action with instanton relations as gauge fixing conditions together with the following Landau type gauge fixing condition for the one form gauge field and the ghost of the shift symmetry: \begin{equation} \partial_\mu\omega^\mu = 0,~~~~~\partial_\mu\widetilde{C}^\mu = 0. \label{eqn:lgf} \end{equation} We introduce a set of antighost fields $\lambda$, $\chi_{\mu a}$, $\overline{\eta}$ and $\overline{C}$, and Lagrange multipliers, $\widetilde{\pi}$, $\pi_{\mu a}$, $\rho$ and $\pi$. These fields obey the standard BRST subalgebra \begin{equation} \begin{array}{rclcrcl} s\lambda&=&\widetilde{\pi}, &\qquad& s\widetilde{\pi}&=&0, \\ s\chi_{\mu a}&=&\pi_{\mu a}, &\qquad& s\pi_{\mu a}&=&0, \\ s\overline{\eta}&=&\rho, &\qquad& s\rho&=&0, \\ s\overline{C}&=&\pi, &\qquad&s\pi&=&0, \label{eqn:ag-brs} \end{array} \end{equation} where anti-self-dual field $\chi_{\mu a}$ obeys the condition ${\epsilon_\mu}^\nu{\epsilon_a}^b\chi_{\nu b}=-\chi_{\mu a}$ and $\pi_{\mu a}$ also obeys the similar condition. We then obtain BRST invariant gauge fixed action \begin{eqnarray} S_{\mbox{\scriptsize{g-f}}} &=& S_0+s\int d^2x\Bigg\{ +\lambda\Big(\frac{1}{2}\epsilon^{\mu\nu}F_{\mu\nu} -|\phi|^2-\beta\widetilde{\pi}\Big) -\chi_{\mu a}\Big((D^\mu\phi)^{a(-)}-\alpha\pi^{\mu a}\Big) \nonumber \\ & &\qquad\qquad\qquad\quad +\overline{\eta}\partial_\mu\widetilde{C}^\mu+\overline{C} \partial_\mu\omega^\mu\Bigg\} \nonumber \\ &=&S_0+\int d^2x\Bigg\{\pi(\partial_\mu\omega^\mu) +\rho(\partial_\mu\widetilde{C}^\mu) \nonumber \\ & &\qquad\qquad\qquad -\pi_{\mu a}\Big((D^\mu\phi)^{a(-)} -\alpha\pi^{\mu a}\Big) +\widetilde{\pi}\Big(\frac{1}{2} \epsilon^{\mu\nu}F_{\mu\nu} -|\phi|^2-\beta\widetilde{\pi}\Big) \nonumber \\ & &\qquad\qquad\qquad +(\partial_\mu\overline{C})(\partial^\mu C+\widetilde{C}^\mu) -\chi_{\mu a}(D^\mu\widetilde{C})^a -\lambda\epsilon^{\mu\nu}\partial_\mu\widetilde{C}_\nu -\partial_\mu\overline{\eta}\partial^\mu\eta \nonumber \\ & &\qquad\qquad\qquad +2\chi_{\mu a}\epsilon^{ab}(D_\mu\phi)_bC -2\chi_{\mu a}\epsilon^{ab}\widetilde{C}^\mu\phi_b -2\lambda\phi_a\widetilde{C}^a\Bigg\}, \label{eqn:bs0} \end{eqnarray} where $\alpha$ and $\beta$ are arbitrary parameters. In the second line of (\ref{eqn:bs0}) BRST transformations of the component fields from (\ref{eqn:ag-brs}), (\ref{eqn:gecu}) and (\ref{eqn:geBi}) are used. \subsubsection{Twisted $N=2$ Super Yang-Mills Action with Dirac- K$\ddot{\hbox{a}}$hler Fermions} We now modify the gauge fixed action $S_{\mbox{\scriptsize{g-f}}}$ of (\ref{eqn:bs0}) by adding BRST exact terms and making the action Hermite and taking the particular choice of the parameters $\alpha$ and $\beta$ \begin{eqnarray} S=\int d^2x &\Big(&+\frac{1}{2}F_{\mu\nu}F^{\mu\nu} + (D_\mu\phi)_a(D^\mu\phi)^a + |\phi|^4 \nonumber \\ &{}& +i\rho\partial_\mu\widetilde{C}^\mu -i\lambda\epsilon^{\mu\nu}\partial_\mu\widetilde{C}_\nu \nonumber \\ &{}& -i\chi_{\mu a}(D^\mu\widetilde{C})^a +\partial_\mu\overline{\eta}\partial^\mu\eta \nonumber \\ &{}& -2i\rho\epsilon^{ab}\phi_a\widetilde{C}_b -2i\lambda\phi^a\widetilde{C}_a -2i\chi_{\mu a}\epsilon^{ab}\widetilde{C}^\mu\phi_b \nonumber \\ &{}& -\frac{i}{4}\epsilon^{\mu\nu} \chi_{\mu a}{\chi_\nu}^a\eta +2i\overline{\eta}\epsilon_{ab}\widetilde{C}^a\widetilde{C}^b +4\overline{\eta}\eta|\phi|^2 \Big). \label{eqn:wa} \end{eqnarray} It is easy to see that the kinetic terms of $\phi_a$, $\widetilde{C}_\mu$, $\rho$, $\lambda$, $\chi_{\mu a}$ and $\widetilde{C}_a$ are nondegenerate while that of $\omega_\mu$ is degenerate. Indeed this action is invariant under the following $SO(2)$ gauge transformations with a gauge parameter $v$, \begin{equation} \begin{array}{rcl} \delta_{\hbox{\tiny gauge}}\omega_\mu &=& \partial_\mu v, \\ \delta_{\hbox{\tiny gauge}}(\phi_a, \widetilde{C}_a, \chi_{\mu a}) &=& 2v{\epsilon_a}^b(\phi_b, \widetilde{C}_b, \chi_{\mu b}), \\ \delta_{\hbox{\tiny gauge}}(\widetilde{C}_\mu, \eta, \lambda, \rho, \overline{\eta}) &=& 0. \\ \end{array} \label{eqn:wgt} \end{equation} We claim that this action possesses BRST symmetry and $N=2$ twisted super algebra up to the above gauge symmetry. The algebra will be summarized for fermionic family of scalar BRST generator $s$ and the vector and pseudo-scalar counterpart $s_\mu$ and $\widetilde{s}$ as follows: \begin{eqnarray} s^2&=&i{\delta_{\hbox{\tiny gauge}}}_{\eta}, \label{eqn:ss}\\ \{ s, s_\mu \} &=& 2i\partial_\mu - 2i{\delta_{\hbox{\tiny gauge}}}_{\omega_{\mu}}, \label{eqn:ssm} \\ \{ s_\mu, s_\nu \} &=& -2ig_{\mu\nu}{\delta_{\hbox{\tiny gauge}}}_{\overline{\eta}}, \label{eqn:smsm}\\ \{ \widetilde{s}, s_\mu \} &=& -2i{\epsilon_\mu}^\nu\partial_\nu + 2i{\delta_{\hbox{\tiny gauge}}}_{{\epsilon_\mu}^\nu\omega_{\nu}}, \label{eq:stsm} \\ \{ \widetilde{s}, \widetilde{s} \} &=& 2i{\delta_{\hbox{\tiny gauge}}}_{\eta}, \label{eq:stst}\\ \{ \widetilde{s}, s \} &=& 0. \label{eqn:sts} \end{eqnarray} The more detailed confirmation of the component wise expressions for the twisted $N=2$ super algebra is given in (\cite{Kawamoto-Tsukioka}). In this formulation we find out very interesting and nontrivial correspondence. We point out that the fermionic fields corresponding to ghosts and Lagrange multipliers can be interpreted as Dirac-K$\ddot{\hbox{a}}$hler fermion fields and thus the twisting procedure is nothing but the Dirac-K$\ddot{\hbox{a}}$hler interpretation of fermionic fields appearing in the quantization procedure. The kinetic terms of multiplet $(\rho, \widetilde{C}_\mu, \lambda)$ and $(\widetilde{C}_a, \chi_{\mu a})$ in the action (\ref{eqn:wa}) can be expressed as \begin{eqnarray} S &=& \int d^2x(i\rho\partial_\mu\widetilde{C}^\mu-i\lambda\epsilon^{\mu\nu} \partial_\mu\widetilde{C}_\nu -i\chi^{\mu a}\partial_\mu\widetilde{C}_a) \nonumber \\ &=& \int d^2x\hbox{Tr}\Big(i\psi^\dagger\gamma^\mu\partial_\mu\psi + i\chi^\dagger\gamma^\mu\partial_\mu\chi\Big), \end{eqnarray} where the Dirac-K$\ddot{\hbox{a}}$hler fields $\psi$ and $\chi$ are given by \begin{eqnarray} \psi &=& \frac{1}{2}(\rho+\widetilde{C}_\mu\gamma^\mu-\lambda\gamma_5), \nonumber \\ \chi &=& \frac{1}{2}(-\widetilde{C}_{a=1} + \chi_{\mu a=1}\gamma^\mu - \widetilde{C}_{a=2}\gamma_5). \label{eqn:dkf} \end{eqnarray} The final expression of the twisted $N=2$ super Yang-Mills action with Dirac-K$\ddot{\hbox{a}}$hler fermions is \begin{eqnarray} S=\int d^2x\Big(&+&\frac{1}{2}F_{\mu\nu}F^{\mu\nu} + (D_\mu\phi)_a(D_\mu\phi)^a + |\phi|^4 \nonumber \\ &+&\frac{1}{2}\partial_\mu A\partial^\mu A - \frac{1}{2}\partial_\mu B\partial^\mu B \nonumber \\ &+& \hbox{Tr}(i\psi^\dagger\gamma^\mu\partial_\mu\psi) \nonumber \\ &+& \hbox{Tr}(i\chi^\dagger\gamma^\mu\partial_\mu\chi) + 2i\omega_\mu\hbox{Tr}(\chi^\dagger\gamma^\mu\chi\gamma_5) \nonumber \\ &-& 4i\phi_1\hbox{Tr}(\psi^\dagger\gamma_5\chi) + 4i\phi_2\hbox{Tr}(\psi^\dagger\gamma_5\chi\gamma_5) \nonumber \\ &-& i\sqrt{2}A\hbox{Tr}(\chi^\dagger\gamma_5 \chi) +i\sqrt{2}B\hbox{Tr}(\chi^\dagger\chi\gamma_5 ) \nonumber \\ &+& 2(A^2-B^2)|\phi|^2 \Big), \label{eqn:dk} \end{eqnarray} where we denote $\eta\equiv \frac{2}{\sqrt{2}}(A+B)$ and $\overline{\eta}\equiv \frac{1}{2\sqrt{2}}(A-B)$. It should be noted that the gauge transformation on Dirac-K$\ddot{\hbox{a}}$hler field $\chi$ can be recognized as $SO(2)$ flavor group. As we have seen in the formulation, the fermionic fields appearing in the quantization procedure such as ghosts and lagrange multiplier turns into the Dirac-K$\ddot{\hbox{a}}$hler matter fermion. It would be important to confirm algebraically that the Dirac-K$\ddot{\hbox{a}}$hler fermions transform as spinor fields and posses half integer spin unlike the ghost fields. Usually $N=2$ super algebra includes the generator $R$ corresponding to a conserved current $R_\mu$ associated with $R$-symmetry. The procedure of the topological twist is related with the redefinition of the stress tensor $T_{\mu\nu}$. We can define new stress tensor $T'_{\mu\nu}$ by the following relation without spoiling current conservation law of $R_\mu$, \begin{equation} T'_{\mu\nu} = T_{\mu\nu} + \epsilon_{\mu\rho}\partial^\rho R_\nu + \epsilon_{\nu\rho}\partial^\rho R_\mu. \end{equation} This modification of the stress tensor leads to a redefinition of the Lorentz rotation generator $J$ \begin{equation} J'=J+R. \label{eqn:nj} \end{equation} This rotation group is interpreted as the diagonal subgroup of $SO(2)\times SO(2)_I$. In other words the topological twist is essentially achieved by identifying spinor and isospinor indices. In the present model we can identify the transformation of the $R$-symmetry is the transformation of the flavor symmetry for the Dirac-K$\ddot{\hbox{a}}$hler fermion fields \begin{equation} \delta_{\scriptsize R}\psi = \psi\Big(\frac{1}{2}\gamma_5\Big), ~~~~~~~~ \delta_{\scriptsize R}\chi = \chi\Big(\frac{1}{2}\gamma_5\Big). \end{equation} Then the Lorentz transformation induced by the generator $J'$ is \begin{equation} \delta_{J'}\psi = -\frac{1}{2}[\gamma_5,\psi], \end{equation} while the Lorentz transformation of $J$ is \begin{equation} \delta_{J}\psi = -\frac{1}{2}\gamma_5\psi. \end{equation} We can obtain the same relations for $\chi$. This implies that Dirac-K$\ddot{\hbox{a}}$hler fermion fields exactly transform as spinors and carry spin one half. \subsection{Weinberg-Salam Model from Generalized Gauge Theory} Connes has pointed out that the Weinberg-Salam model can be formulated as a particular case of the noncommutative geometry formulation of a gauge theory\cite{Connes}. Roughly speaking his idea is the following. Consider a manifold which is composed of a direct product of discrete two points and four dimensional flat space and define a connection and differential operator on this manifold. Due to the discrete nature of the two points the differential operator can be represented by two by two matrix. Thus the connection or equivalently the gauge field is now represented by two by two matrix and thus possesses diagonal and off-diagonal components. Then the weak and electromagnetic one form gauge fields are assigned to the diagonal component while the zero form Higgs fields are assigned to the off-diagonal components. Then the spontaneously broken Weinberg-Salam model comes out naturally from the pure Yang-Mills action by taking the group $SU(2)\times U(1)$\cite{Connes}\cite{Coq}. \subsubsection{Generalized Gauge Theory with Dirac-K$\ddot{\hbox{a}}$hler Fermions} In this subsection we show that the two by two matrix representation of the gauge fields are easily accommodated by the quaternions of the generalized gauge field since the quaternion algebra can be represented by two by two matrices. We can, however, point out that our generalized gauge theory is more general formulation as noncommutative geometry since it includes not only zero and one form of gauge fields but also all the possible form degrees of gauge fields\cite{KTU}. In formulating generalized Yang-Mills action we need to define the inner product to construct the action. In the sense of the generalized topological Yang-Mills theory the inner product has been defined by the naive wedge product. Here we intend to incorporate fermionic form degrees of freedom in accordance with the generalized gauge field. We then formulate the fermionic form degrees by Dirac-K$\ddot{\hbox{a}}$hler fermion formulation to generate matter fermions. We thus need to define a new inner product to accommodate both the generalized gauge fields and Dirac-K$\ddot{\hbox{a}}$hler fields. We keep to use the generalized gauge field and parameter as defined above in (\ref{eqn:def-gf}) and (\ref{eqn:def-gp}). In addition we introduce fermionic matter generalized field \begin{eqnarray} \bm{$\psi$} &=& \mbox{\bf{1}}\psi + \mbox{\bf{i}}\hat{\psi} \in \Lambda_- \nonumber \\ \overline{\bm{$\psi$}} &=& \mbox{\bf{j}}\hat{\overline{\psi}} + \mbox{\bf{k}}\overline{\psi} \in \Lambda_+, \end{eqnarray} where $\psi$ and $\overline{\psi}$ fermionic odd forms while $\hat{\psi}$ and $\hat{\overline{\psi}}$ are fermionic even forms. We now extend the generalized differential operator ${\cal{Q}} = {\mbox{\bf{j}}}d$ into \begin{equation} {\cal D} = \mbox{\bf{j}}d + \mbox{\bf{k}}m, \label{eqn:gy-ed} \end{equation} where $m$ is a bosonic constant matrix. In this subsection we consider the formulation only in four dimensions. It is straightforward to generalize the formulation into arbitrary even dimensions. This differential operator has the following grading structure: \begin{equation} {\cal DA} = \{ \ {\cal D}, \ {\cal A} \ \}, \quad {\cal A} \in \Lambda_{-}, \end{equation} \begin{equation} {\cal DV} = [ \ {\cal D}, \ {\cal V} \ ], \quad {\cal V} \in \Lambda_{+}. \end{equation} The graded Leibniz rule can be easily checked, but the nilpotency is not satisfied in general: ${\cal{D}}^2 = -{\mbox{\bf{1}}}m^2$. The gauge transformation of the generalized curvature has extra factor due to the non-nilpotency of the differential operator \begin{eqnarray} \delta {\cal{F}} &=& [ \ {\cal{F}}, \ {\cal{V}} \ ] + [ \ {\cal{D}}^2, \ {\cal{V}} \ ] \nonumber \\ &=& [ \ {\cal{F}}, \ {\cal{V}} \ ] - [ \ m^2, \ {\cal{V}} \ ]. \label{eqn:delcurv} \end{eqnarray} In order to define generalized Yang-Mills action with Dirac-K$\ddot{\hbox{a}}$hler fermions as matter we introduce the notion of Clifford product or simply cup product $\vee$ which was introduced by K\"{a}hler and should be differentiated from the wedge product $\wedge$\cite{DK}\cite{BJ}. We consider an element $u$ which is the direct sum of forms \begin{eqnarray} u= \sum^m_{p=0}\frac{1}{p!}u_{\mu_1\cdots\mu_p} dx^{\mu_1}\wedge\cdots\wedge dx^{\mu_p}, \label{eqn:gy-gf} \end{eqnarray} where $m\le d$ with $d$ as spacetime dimension. We then define the linear operator $e_{\mu}$, \begin{eqnarray} e_\mu u= \sum^{m-1}_{p=0}\frac{1}{p!}u_{\mu\mu_1\cdots\mu_p} dx^{\mu_1}\wedge\cdots\wedge dx^{\mu_p}, \label{eqn:gy-cop} \end{eqnarray} which is understood as a differentiation of the polynomial of differential form with respect to $dx^\mu$. In particular it has the role of contracting operator as \begin{eqnarray} e_\mu dx^{\alpha_1}&\wedge& dx^{\alpha_2}\wedge dx^{\alpha_3}\wedge dx^{\alpha_4} \wedge\cdots \nonumber \\ &=& g_\mu^{\alpha_1}dx^{\alpha_2}\wedge dx^{\alpha_3}\wedge dx^{\alpha_4} \wedge\cdots - g_\mu^{\alpha_2}dx^{\alpha_1}\wedge dx^{\alpha_3} \wedge dx^{\alpha_4} \wedge\cdots \nonumber \\ &+& g_\mu^{\alpha_3}dx^{\alpha_1}\wedge dx^{\alpha_2} \wedge dx^{\alpha_4} \wedge\cdots - g_\mu^{\alpha_4}dx^{\alpha_1}\wedge dx^{\alpha_2} \wedge dx^{\alpha_3} \wedge\cdots + \cdots . \label{eqn:gy-copex} \end{eqnarray} We further define sign operators $\eta$ and $\zeta$ which generate the following sign factors: \begin{eqnarray} \zeta u_p &=& \zeta_p u_p \equiv (-1)^{\frac{p(p-1)}{2}} u_p \nonumber \\ \eta u_p &=& (-1)^p u_p, \label{eqn:gy-sf} \end{eqnarray} where $u_p$ is a p-form variable. We can now define the cup (Clifford) product of $u$ and $v$ \begin{eqnarray} u\vee v= \sum^m_{p=0} \frac{1}{p!} \zeta_p \{ \eta^p(e_{\mu_1}\cdots e_{\mu_p} u) \} \wedge e^{\mu_1}\cdots e^{\mu_p} v, \label{eqn:gy-cp} \end{eqnarray} where $v$ has the similar form as $u$ in (\ref{eqn:gy-gf}). The sign factor $\zeta_p$ and the operator $\eta^p$ are necessary to confirm the associativity of the cup product \begin{eqnarray} (u\vee v)\vee w = u\vee (v\vee w) . \label{eqn:gy-associativity} \end{eqnarray} We introduce the generalized Yang-Mills action \begin{equation} S_G = \int \hbox{Tr}[\overline{\bm{$\psi$}}\vee({\cal D} +{\cal A}){\bm{$\psi$}} + {\cal{F}}\vee {\cal{F}}]_{\hbox{{\bf 1}}} * 1 \label{eqn:gy-gga} \end{equation} where $*$ is Hodge star operator and in particular $* 1 = \sqrt{g}\frac{1}{4!}\epsilon_{\alpha_1\alpha_2\alpha_3\alpha_4} dx^{\alpha_1}\wedge dx^{\alpha_2}\wedge dx^{\alpha_3}\wedge dx^{\alpha_4} $ in four dimensions with $g$ as the metric determinant. The simbol ${\bf 1}$ denotes to pick up the coefficient of ${\bf 1}$ for the quaternion expansion in the trace. We then define the generalized gauge parameter \begin{eqnarray} { \cal{V}}= \hbox{{\bf 1}} v, \label{eqn:gy-ggp} \end{eqnarray} where $v$ is 0-form gauge parameter and thus ${\cal{V}} \in \Lambda_+$. This has a clear contrast with the generalized gauge parameter of generalized Chern-Simons formulation where all the degrees of differential forms were introduced. We now claim that the generalized Yang-Mills action with Dirac-K$\ddot{\hbox{a}}$hler matter fermion is gauge invariant under the following gauge transformation: \begin{eqnarray} \delta {\cal A} &=& [{\cal{D}}+{\cal{A}},{\cal{V}}]_{\vee}, \nonumber \\ \delta {\bm{$\psi$}} &=& - {\cal V}\vee {\cal \psi}, \nonumber \\ \delta \overline{\bm{$\psi$}} &=& \overline {\cal \psi}\vee {\cal V}, \label{ym-ggt} \end{eqnarray} where we have used the notation $[\cal{A},\cal{B}]_\vee = \cal{A}\vee\cal{B}-\cal{B}\vee\cal{A}$. Here we should mention that in the quaternion algebra we should choose the sign choice $(\epsilon_1,\epsilon_2)=(1,1)$ to prove the generalized gauge invariance of the generalized Yang-Mills action (\ref{eqn:gy-gga}). \subsubsection{Weinberg-Salam Model from Generalized Gauge Theory as Non-Commutative Geometry Formulation} In the formulation of previous subsection we have introduced all the degrees of differential forms but only zero form for the generalized gauge parameter. The reason why we have introduced only zero form gauge parameter is that the action is not gauge invariant under the higher form gauge parameters. Hereafter we introduce only zero form and one form gauge fields for the generalized gauge field and do not consider the Dirac-K$\ddot{\hbox{a}}$hler fermionic matter for simplicity. We take the algebra of super group $SU(2|1)$ as the graded Lie algebra. $SU(2|1)$ generators can be represented by $3\times3$ matrices~\cite{Coq} $$ T_{i}=\left( \begin{array}{@{\,}cc|c@{\,}} & \mbox{ \hspace{-7.6mm} \raisebox{-2.5mm}[0pt][0pt] { $ \displaystyle \frac{\sigma_{i}}{2} $ } } & 0 \\ & & 0 \\ \hline 0 & 0 & 0 \end{array} \right), \quad Y =\left( \begin{array}{@{\,}cc|c@{\,}} 1 & 0 & 0 \\ 0 & 1 & 0 \\ \hline 0 & 0 & 2 \end{array} \right), $$ $$ \Sigma_{+}=\left( \begin{array}{@{\,}cc|c@{\,}} 0 & 0 & 1 \\ 0 & 0 & 0 \\ \hline 0 & 0 & 0 \end{array} \right), \ \Sigma_{-}=\left( \begin{array}{@{\,}cc|c@{\,}} 0 & 0 & 0 \\ 0 & 0 & 1 \\ \hline 0 & 0 & 0 \end{array} \right), \ \Sigma'_{+}=\left( \begin{array}{@{\,}cc|c@{\,}} 0 & 0 & 0 \\ 0 & 0 & 0 \\ \hline 0 & 1 & 0 \end{array} \right), \ \Sigma'_{-}=\left( \begin{array}{@{\,}cc|c@{\,}} 0 & 0 & 0 \\ 0 & 0 & 0 \\ \hline 1 & 0 & 0 \end{array} \right), \ $$ where $\sigma_{i}$'s are Pauli matrices. They satisfy the following graded Lie algebra: $$ [ \ T_{i}, \ T_{j} \ ]=i\epsilon_{ijk}T_{k}, \quad [ \ Y, \ T_{i} \ ]=0, $$ $$ [ \ T_{\pm}, \ \Sigma_{\pm} \ ]=0, \quad [ \ T_{\pm}, \ \Sigma_{\mp} \ ]=\frac{1}{\sqrt{2}}\Sigma_{\pm}, $$ $$ [ \ T_{\pm}, \ \Sigma'_{\pm} \ ]=0, \quad [ \ T_{\pm}, \ \Sigma'_{\mp} \ ]=-\frac{1}{\sqrt{2}}\Sigma'_{\pm}, $$ $$ [ \ T_{3}, \ \Sigma_{\pm} \ ]=\pm\frac{1}{2}\Sigma_{\pm}, \quad [ \ T_{3}, \ \Sigma'_{\pm} \ ]=\pm\frac{1}{2}\Sigma'_{\pm}, $$ $$ [ \ Y, \ \Sigma_{\pm} \ ]=-\Sigma_{\pm}, \quad [ \ Y, \ \Sigma'_{\pm} \ ]=\Sigma'_{\pm}, $$ $$ \{ \ \Sigma_{\pm}, \ \Sigma_{\pm} \ \} = \{ \ \Sigma_{\pm}, \ \Sigma_{\mp} \ \}=0, \quad \{ \ \Sigma'_{\pm}, \ \Sigma'_{\pm} \ \} = \{ \ \Sigma'_{\pm}, \ \Sigma'_{\mp} \ \}=0, $$ $$ \{ \ \Sigma_{\pm}, \ \Sigma'_{\pm} \ \}=\sqrt{2}T_{\pm}, \quad \{ \ \Sigma_{\pm}, \ \Sigma'_{\mp} \ \}=\pm T_{3}+\frac{1}{2}Y, $$ with $T_{\pm}=\frac{1}{\sqrt{2}}(T_{1}\pm iT_{2})$. $T_{3}, Y$ correspond to a generator of a weak isospin and a weak hypercharge, respectively. This algebra contains $SU(2)\times U(1)_{Y}$ in the even graded parts $T_{i}, Y$, and $SU(2)$ doublet in the odd graded parts $\Sigma_{\pm}, \Sigma'_{\pm}$ whose subscripts $\pm$ correspond to the generators with eigenvalues $\pm\frac{1}{2}$ of the generator $T_{3}$. Indeed Higgs doublet which we have introduced as Lie algebra valued gauge fields corresponds to these odd parts. It is interesting to note that the super symmetric algebra of $SU(2|1)$ can be accommodated in the generalized gauge theory even without fermionic fields. The gauge field is now expanded by corresponding fields of the generators \begin{equation} {\cal{A}}={\mbox{\bf{j}}} \Big(iA^{i}T_{i}+\frac{i}{2\sqrt{3}}BY\Big) +{\mbox{\bf{k}}} \Big(\frac{i}{\sqrt{2}}(\phi_{0}\Sigma_{+}+\phi_{+}\Sigma_{-} +\phi_{0}^{*}\Sigma'_{-} +\phi_{+}^{*}\Sigma'_{+})\Big), \end{equation} where $A^{i}, B$ and $\phi_{0}, \phi_{+}$ are real $SU(2)\times U(1)_{Y}$ gauge fields and complex Higgs scalar fields respectively. We can choose a particular form of the constant matrix $m$ of the generalized differential operator $m=\frac{i}{\sqrt{2}}(v\Sigma_{+}+v^*\Sigma'_{-})$ which leads \begin{equation} m^2=-\frac{|v|^2}{2}\Big(T_{3}+\frac{1}{2}Y\Big). \label{em} \end{equation} Then eq. (\ref{em}) exactly represents the electro-magnetic charge. As we can see from the gauge transformation of the curvature in (\ref{eqn:delcurv}) the gauge symmetries $v$ spontaneously break down to the special direction which is required from the consistency of the extended differential algebras. The curvature is given with a suitable redefinition of fields ${\cal{A}}\rightarrow g{\cal{A}}$, \begin{equation} {\cal{F}}={\mbox{\bf{1}}} \bigg(g^2{\cal{F}}^{(0)} -\frac{i}{2}g{\cal{F}}_{\mu\nu}^{(2)} dx^{\mu}\wedge dx^{\nu} \bigg) +{\mbox{\bf{i}}}ig{\cal{F}}_{\mu}^{(1)}dx^{\mu}. \end{equation} The kinetic terms of $SU(2)\times U(1)_{Y}$ gauge fields are \begin{equation} {\cal{F}}_{\mu\nu}^{(2)}=F_{\mu\nu}^kT_{k}+\frac{1}{2\sqrt{3}}G_{\mu\nu}Y, \end{equation} where $$ F_{\mu\nu}^{k}=\partial_{\mu}A_{\nu}^{k}-\partial_{\nu}A_{\mu}^{k} -2g\epsilon_{ij}{}^{k}A_{\mu}^{i}A_{\nu}^{j}, $$ $$ G_{\mu\nu}=\partial_{\mu}B_{\nu}-\partial_{\nu}B_{\mu}. $$ The kinetic terms of Higgs fields and interaction terms are \begin{eqnarray} {\cal{F}}_{\mu}^{(1)}=\frac{1}{\sqrt{2}}\Bigg\{ &\bigg(&\partial_{\mu}\phi_{0} +\frac{i}{\sqrt{2}}gW_{\mu}\phi_{+} +\frac{i}{\sqrt{3}}gZ_{\mu} \Big(\phi_{0}+\frac{v}{g}\Big) \bigg)\Sigma_{+} \nonumber \\ + &\bigg(&\partial_{\mu}\phi_{+} +\frac{i}{\sqrt{2}}gW_{\mu}^{\dagger} \Big(\phi_{0}+\frac{v}{g}\Big) -i\frac{\sqrt{3}}{6}gZ_{\mu}\phi_{+} -\frac{i}{2}gA_{\mu}\phi_{+} % \bigg)\Sigma_{-} \nonumber \\ +&{}&{\mbox{h.c.}} \ \Bigg\}, \end{eqnarray} where $$ W_{\mu}=\frac{1}{\sqrt{2}}\Big( A_{\mu}^{1}-iA_{\mu}^{2} \Big), $$ and \begin{eqnarray} Z_{\mu} &=& \frac{\sqrt{3}}{2}A_{\mu}^3-\frac{1}{2}B_{\mu}, \nonumber \\ A_{\mu} &=& \frac{1}{2}A_{\mu}^3+\frac{\sqrt{3}}{2}B_{\mu}. \label{az} \end{eqnarray} These identifications (\ref{az}) fix the Weinberg angle to be $\theta_{W}=\frac{\pi}{6}$ which is an arbitrary parameter in the Weinberg-Salam model. Thus the direction of spontaneous breaking is particularly chosen by the model itself. The Higgs potential term is given by \begin{eqnarray} {\cal{F}}^{(0)}=\frac{1}{2} &&\Bigg\{\bigg(\Big|\phi_{0}+\frac{v}{g}\Big|^2 -\Big|\frac{v}{g}\Big|^2\bigg) \Big(T_{3}+\frac{1}{2}Y\Big) +\bigg(\phi_{0}+\frac{v}{g}\bigg) \phi_{+}^{*}\sqrt{2}T_{+} \nonumber \\ && +\phi_{+}\bigg(\phi_{0}^{*}+\frac{v^{*}}{g}\bigg) \sqrt{2}T_{-} +|\phi_{+}|^2\Big(-T_{3}+\frac{1}{2}Y\Big)\Bigg\}. \label{c0} \end{eqnarray} Then the full expression of the action in Minkowski space-time is given by \begin{eqnarray} S = \int d^4x \Bigg\{ &-& \frac{1}{4}F_{\mu\nu}F^{\mu\nu} - \frac{1}{4}Z_{\mu\nu}Z^{\mu\nu} \nonumber \\ &-&\frac{1}{2}(D^{\mu\dagger}W^{\nu\dagger}-D^{\nu\dagger}W^{\mu\dagger}) (D_{\mu}W_{\nu}-D_{\nu}W_{\mu}) \nonumber \\ &-&2ig\Big(\frac{\sqrt{3}}{2}Z_{\mu\nu} +\frac{1}{2}F_{\mu\nu}\Big)W^{\mu}W^{\nu\dagger} \nonumber \\ &+&2g^2\Big(|W_{\mu}W^{\mu}|^2-(W_{\mu}W^{\mu\dagger})^2\Big) \nonumber \\ &+&\Big|{\partial}_{\mu}\phi_{0}+\frac{i}{\sqrt{2}}gW_{\mu}\phi_{+} +\frac{i}{\sqrt{3}}Z_{\mu} \Big(\phi_{0}+\frac{v}{g}\Big)\Big|^2 \nonumber \\ &+&\Big|{\partial}_{\mu}\phi_{+} +\frac{i}{\sqrt{2}}gW_{\mu}^{\dagger} \Big(\phi_{0}+\frac{v}{g}\Big) -i\frac{\sqrt{3}}{6}gZ_{\mu}\phi_{+} -\frac{i}{2}gA_{\mu}\phi_{+} \Big|^2 \nonumber \\ &-&\frac{g^2}{2}\bigg( \Big|\phi_{0}+\frac{v}{g}\Big|^2 + |\phi_{+}|^2 \bigg)^2 \nonumber \\ &+&\frac{3}{4}g^2\Big|\frac{v}{g}\Big|^2(1-2y) \bigg( \Big|\phi_{0}+\frac{v}{g}\Big|^2+|\phi_{+}|^2 \bigg) \nonumber \\ &-&\frac{3}{8}g^2\Big|\frac{v}{g}\Big|^2(1-2y)^2 \Bigg\}, \end{eqnarray} where \begin{eqnarray*} F_{\mu\nu} &=& \partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu}, \\ Z_{\mu\nu} &=& \partial_{\mu}Z_{\nu}-\partial_{\nu}Z_{\mu}, \\ D_{\mu} &=& \partial_{\mu}+i2g\Big(\frac{\sqrt{3}}{2}Z_{\mu} +\frac{1}{2}A_{\mu} \Big). \end{eqnarray*} Higgs potential term can be minimized by shifting the field with $y=-\frac{1}{6}$. This leads to the ratio of the mass of physical Higgs $M_{\phi}$ and weak boson $M_{W}$ to $\frac{M_{\phi}}{M_{W}}=\sqrt{2}$. The predictions from this scheme agree with noncommutative geometric models~\cite{Connes}\cite{Coq}. This model has fewer parameters than the usual Weinberg-Salam model and has predictive power on the Higgs mass. \section{Possible Scenario and Conjectures for Unified Model on the Lattice} I would like to discuss a possible scenario and then propose conjectures of unified model on the lattice based on the analyses of previous sections and suggestive known results. Constructive and reguralized quantum theory of gravity will be given by dynamically triangulated simplicial lattice manifold. In two dimensional quantum gravity we have enough evidences that this is the fact and we have the analytical and numerical formulations to describe it. As we have seen in the three dimensional example of Chern-Simons gravity the lattice manifold may be interpreted as spin graph where links carry the quantum numbers of angular momentum. Two dimensional topological gravity has similar nature as the Chern-Simons gravity where links may carry conformal spin quantum numbers\cite{Verlinde-Verlinde}. In four dimensional lattice gravity formulations there are similar evidences\cite{Ooguri}\cite{FDgravity}\cite{Smolin}. Gauge theory formulation of the lattice gravity and matter will be formulated by the generalized gauge theory. In other words the generalized gauge theory formulated by differential forms may play a crucial role in formulating a lattice gravity since the diffeomorphism invariance is automatic in the formulation of differential forms. Furthermore differential forms and simplexes have one to one correspondence and thus the fields on the forms can naturally be put on the corresponding simplexes. Quaternion algebra may play a crucial role in the formulation of the lattice gravity formulated by the generalized gauge theory. In other words even forms, odd forms, fermions and bosons are the fundamental algebraic ingredients which are nicely classified by the quaternion algebra. Furthermore the graded Lie algebra will be another fundamental algebra which governs the gauge structure of the generalized gauge theory, in particular, the super symmetry. Thus the semi-direct product of the graded Lie algebra and the quaternion algebra is the fundamental structure of the generalized gauge theory and may play an important role in formulating lattice gravity as a gauge theory. Fermionic matter will be generated from the ghosts of quantization by the twist. The twisting procedure is essentially related with the Dirac K$\ddot{\hbox{a}}$hler fermion formulation as we have shown in the quantization of topological Yang-Mills action. The fermionic ghosts appear as fermionic differential forms and can be put on the simplicial lattice similar as the bosonic fields. We need to understand this twisting mechanism from lattice point of view. As we have seen in the quantization of the generalized Chern-Simons actions we needed to introduce infinite number of ghosts due to the infinite reducibility of the system\cite{KSTU}. It is interesting to consider possible physical aspects of the introduction of an infinite number of the ghost fields. An immediate consequence is a democracy of ghosts and classical fields, {\it i.e.}, the classical fields are simply the zero ghost number sector among infinitely many ghost fields and thus the classical gauge fields and ghost fields have no essential difference in the minimal action. Furthermore fermionic and bosonic gauge fields are treated in an equal footing and the series of infinite ghosts originated from the classical fermionic and bosonic fields are complementary. In other words if we only introduce bosonic classical fields in the starting action we need to introduce fermionic fields with odd integer ghost number and bosonic fields with even integer ghost number as we have shown in section 5. If we introduce the classical fermionic gauge fields, the odd and even nature should be reversed for the ghost numbers when introducing the corresponding ghost fields to the fermionic gauge fields. It seems to mean that even the fermionic and bosonic fields have no essential difference in the generalized Chern-Simons theory. In other words fermionic fields, bosonic fields, classical fields and ghost fields are mutually inter-related via the quantization procedure. Super symmetry is the natural consequence of topological symmetry with partial gauge fixing of instanton conditions. Thus the introduction of super symmetry is realized via quantization procedure. In the generalized gauge theory graded Lie algebra plays an important role leading to the super symmetry. The lessons from the lattice Chern-Simons gravity of the section 4 tell us that the drei bein and spin connection can be simply put on the original and dual lattice, respectively. This means that the fields of differential forms from the generalized gauge theory will be put on both of the original and dual lattice. The standard model including Weinberg-Salam model will be realized by the generalized gauge theory via non-commutative geometry mechanism. In particular the quaternion algebra plays an essential role here again to formulate the gauge theory of all the differential forms. It should be noted that Connes's non-commutative geometry formulation of Weinberg-Salam model is a particular example of our generalized gauge theory formulation. Connes has speculated that the discrete two points which he introduced in the non-commutative geometry formulation can be interpreted as the discrete distance of the space time structure which is physically related with the Higgs vacuum expectation value\cite{Connes}. In our generalized gauge theory the quaternion algebra is playing a fundamental role and should have some physical or geometrical interpretation. I would like to give conjectures on the possible interpretation of the quaternion and the fermion-boson correspondence following to the suggestive examples of analyses given in this summary. \begin{center} {\bf Conjectures} \end{center} The generalized gauge theory will be formulated on the simplicial lattice manifold and may play a fundamental role in formulating a unified model on the lattice. Quaternion algebra is responsible for differentiating even simplexes, odd simplexes corresponding to even forms and odd forms, and original simplicial lattice and the dual simplicial lattice reflecting boson and fermion nature. In particular discrete two points in terms of non-commutative geometry would correspond to the original lattice and dual lattice. Matter fermions are generated from ghosts via twisting mechanism which is nothing but the Dirac K$\ddot{\hbox{a}}$hler fermion formulation. Bosons and fermions are dual to each other in the lattice simlex sense. In other words bosons are sitting on the original lattice while fermions are sitting on the dual lattice and {\it vice versa}. Chiral fermion problem will be understood via Dirac K$\ddot{\hbox{a}}$hler fermion formulation. This duality relation would be analogous to the Ising model where the disorder variable defined on the dual lattice is related to the fermion field as shown in the subsection 2.1. Strong coupling and weak coupling duality of Krammers and Wannier type is working here and the fermion will be a soliton on the dual lattice. The origin of the Seiberg and Witten formulation\cite{Seiberg-Witten} will be related to the duality of original and dual lattice. Due to the complete duality of the original lattice and the dual lattice where fermions and bosons are located on the mutually dual lattices the super symmetry is the natural consequence for the whole system. The super symmetry transformation and the gauge transformation are the change of the fields from one simplex to the other on the original lattice and the dual lattice. This is analogous to the lattice gauge theory where even sites and odd sites exchange corresponds to the chiral transformation. Instanton will play an essential role in this bosonization mechanism and also in the breaking of the higher super symmetry or the topological symmetry. The phenomenological results like the group structure of the standard model and the number of the generations mentioned in the introduction will be understood from this point of view. \begin{center} {\bf Questions} \end{center} There are still many fundamental questions. In this summary we have mainly discussed topological field theories which do not include dynamical degrees of freedom. Then the question arises: \\ \lq\lq How do the dynamical degrees of freedom appear starting from the theory with higher symmetry such as $N=2$ super symmetry and topological symmetry ?'' \\ We can guess that the classical solution like instanton may play an important role to resolve the higher symmetry into lower symmetry but the mechanism is not yet clear. The topological symmetry or $N=2$ super symmetry may play a similar role as the kinetic term in the lagrangian formulation of field theory. The field theories keeping higher symmetry such as topological symmetry or higher super symmetry may not yet have genuine non-perturbative interactions which break the symmetry dynamically. There are other questions concerning to the number of dimensions of our space-time and Minkowskian nature. String approach may be able to understand this question\cite{Kawai1}. I believe that the string approach and our approach is not incompatible but rather complementary. \vskip 1cm \noindent{\Large{\bf Acknowledgements}} \\ I would like to thank my collaborators J.Ambj\o rn, K.N.Anagnostopoulos, T.Ichihara, L.Jensen, H. Kawai, V.A. Kazakov, T. Mogami, H.B.Nielsen, E.Ozawa, Y. Saeki, N.Sato, J.Smit, K.Suehiro, T.Tsukioka, H.Umetsu, Y. Watabiki and K.Yotsuji for fruitful discussions and collaborations .
\section{Introduction} The dynamic behavior of the gas in the inner few pc of the Galaxy is still a matter of debate and speculation. There is an almost disklike structure of neutral gas located around the Galactic Center (GC) \cite{GGWal87,JGGal93} \cite{SGWal86}. This {\it Circum-Nuclear Disk\/} (CND) extends over a radial distance of more than 5\,pc. It has an inner edge at about 2\,pc from the GC. The predominantly molecular gas moves around the GC with a speed of $\sim$100\,km\,s$^{-1}$ corresponding to a Keplarian velocity for an enclosed mass of several 10$^{6}$\,M$_{\odot}$. Within the inner 2\,pc ionized gas has been found \cite{LC83} which has the appearance of a spiral and was therefore named the {\it minispiral\/}. The region within the CND is often referred to as {\it Sgr A West\/}. It can be separated into at least four different components \cite{LAS91} which we will discuss below. The kinematics derived from the \neii\ data shows a possible close link between this ionized material and the molecular gas at the inner rim of the CND. Thus the outer part of the minispiral appears to be the ionized edge of the CND. Often the minispiral was thought to be not only of spiral shape in projection against the celestial sphere, but rather a true one-armed spiral. However, the kinematics derived from radio recombination line observations \cite{RG93} indicated several different structures instead of one single spiral. In this paper we re-examine the \neii\ data and the kinematic information contained in it. First, we introduce the dataset that we use (Sect.\ \ref{sec:data}), and evaluate its usability as a tracer of ionized gas by comparing it with radio recombination lines (Sect.\ \ref{sec:radio}). In Sect.\ \ref{sec:vis}, we describe our visualization technique which proved to be of paramount importance for our analysis. Section \ref{sec:morph} is devoted to a discussion of the morphological details in the image of the minispiral. The model that we derive from the data is presented in Sect.\ \ref{sec:model}, in the subsequent Section, we describe the three dimensional structure of the gas streams, and summarize our conclusions in Sect \ref{sec:concl}. \section{The data\label{sec:data}} We use the \neii ($\lambda 12.8\,\mu$m) line emission observations described in detail by \citeasnoun{LAS91}. The data cube was made available to us by Lacy. The observations were carried out using a mid-infrared cryogenic echelle spectrograph. The spectral dispersion was about 16.5\,km\,s$^{-1}$ per pixel, the spectral resolution being about 2 pixels FWHM. An area of $75^{\prime\prime}$ $\times$ 90$^{\prime\prime}$ was mapped with 2$^{\prime\prime}$ resolution. A \neii\ line map was produced by summing over the channels covering Doppler shifts from $+380$ to $-412$\,km\,s$^{-1}$. This map is shown as a ruled-surface representation in Figure \ref{fig:map}. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure2.eps}} \caption{Ruled-surface representation of the \neii\ line emission. The intensity is given in units of 10$^{-2}$\,erg\,s$^{-1}$\,cm $^{-2}$\,sr$^{-1}$. All velocity channels are summed. \label{fig:map}} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure3.ps}} \caption{Superposition of the 8.3\,GHz data as a contour representation together with the \neii\ line emission in a linear grey scale. Intensities greater than 3\,10$^{-3}$\,erg\,s$^{-1}$\,cm$^{-2}$\,sr$^{-1}$ are set to this value in order to emphasize the features. The point source Sgr A* is indicated with a cross.\label{fig:radio}} \end{figure} \section{The comparison with the radio data\label{sec:radio}} \citeasnoun{RG93} observed the Sgr A West complex in the H92$\alpha$ line at 8.3\,GHz with a resolution of 1$^{\prime\prime}$. As the data of the radio recombination line and the \neii\ emission at 12.8 $\mu$m give both informations about the distribution of the ionized gas, they should show the same features. Figure \ref{fig:radio} shows the superposition of these data. As, at this point, we are not interested in fine details but in much more in the overall distributions, the different resolutions of the two datasets pose no problem. We find that the overall features are the same in both wavelengths indicating that \begin{enumerate} \item there are no depopulation effects affecting the \neii\ data, \item \neii\ line emission is a good tracer for the ionized gas \end{enumerate} \section{Three dimensional visualisation\label{sec:vis}} In the following, we will often use a three dimensional representation of the data using a normal Cartesian coordinate system with R.A. corresponding to the $x$ direction, Dec. to the $y$ direction and the local standard of rest (LSR) velocities to the $z$ direction. To give a better visual impression, we have chosen to illuminate the shown surfaces from the observer's direction. In this representation features are the brighter the closer they are to the observer. The surfaces indicate pre-determined levels of constant intensity. To keep the figures as clear as possible, we give absolute numbers for the axes only in the color representation of Fig.\ \ref{fig:color}. \begin{figure*} \resizebox{\hsize}{!}{\includegraphics{figure1.eps}} \caption{A three dimensional representation of the data cube (radial velocity as a function of position on the sky. The velocities are color coded. In the bottom plane (R.A./Dec.), the projected minispiral is depicted. For an even clearer visualization of the 3D structure of the data cube, we also show it in an animation (\harvardurl{anim1.gif}). \label{fig:color}} \end{figure*} \section{The morphology\label{sec:morph}} \citeasnoun{LAS91} discussed four separate morphological components in the Sgr A West complex: \begin{enumerate} \item the {\it Northern Arm\/} which is the most prominent feature and runs approximately from the radio source Sgr A* (the {\it Galactic Center\/}) to the north; \item the {\it Western Arc\/} which is the bright, almost "vertical" rim in the west of Sgr A*; \item the {\it Eastern Arm\/} which runs from the north east to the east of Sgr A*; and \item the {\it Bar\/} which is situated "below" Sgr A* and and shows a straight shape. \end{enumerate} Applying the above described analysis technique, it becomes evident that at least for the Bar the structure is more complicated than previoulsy thought. Figures \ref{fig:three} to \ref{fig:five} show the data cube rotated by 30$^\circ$ around the $x$- and $z$-axis for three different intensity levels. Figures \ref{fig:five}, \ref{fig:six}, and \ref{fig:seven} give different viewing angles to allow for a full appreciation of the three dimensional distribution. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure4.ps}} \caption{The Sgr A West complex shown in an intensity level of 1\,10$^{-2}$\,erg\,s$^{-1}$\,cm$^{-2}$\,sr$^{-1}$. For the absolute scaling, see Fig.\ \ref{fig:color}.\label{fig:three}} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure5.ps}} \caption{The Sgr A West complex shown in an intensity level of 6\,10$^{-3}$\,erg\,s$^{-1}$\,cm$^{-2}$\,sr$^{-1}$. For the absolute scaling, see Fig.\ \ref{fig:color}.\label{fig:four}} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure6.ps}} \caption{The Sgr A West complex shown in an intensity level of 3\,10$^{-3}$\,erg\,s$^{-1}$\,cm$^{-2}$\,sr$^{-1}$. This level is used for the comparison with the model calculations.For the absolute scaling, see Fig.\ \ref{fig:color}.\label{fig:five}} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure7.ps}} \caption{The Sgr A West complex shown in an intensity level of 3\,10$^{-3}$\,erg\,s$^{-1}$\,cm$^{-2}$\,sr$^{-1}$. The cube is rotated by 120$^\circ$ with respect to the $z$-axis. Under this viewing angle, the two different components of the Bar can be seen particularly well.\label{fig:six}} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure8.ps}} \caption{The same representation of the data cube as in Fig.\ \ref{fig:six}, but rotated by 300$^\circ$ with respect to the $z$-axis. Under this angle, the Western Arc can be seen particularly well as a bright line of emitting material in the foreground.\label{fig:seven}} \end{figure} The following features are marked: \begin{enumerate} \item {\bf The Northern Arm:} It appears as an almost vertical tube where material with the highest densities can be found. If one looks at Figure \ref{fig:four}, it becomes apparent, that part of the high negative velocity end of this feature runs into the area of the Bar. \item {\bf The Western Arc:} It represents the feature of lowest intensity, and as such appears only in the images of the data cube that trace the lowest intensities. There it can be seen as a bent tube with a small velocity gradient. \item {\bf The Eastern Arm:} It appears to consist of two components. A vertical {\it finger\/} of high intensity and a large {\it ribbon\/} which extends to the east of Sgr A*. \item {\bf The Bar:} As the polarization measurements of \citeasnoun{AGSal91} have already shown, this is the most complicated feature in the region. Looking at Figs. \ref{fig:three} and \ref{fig:four} it is clear that there are two distinctly different components which we will call {\it Bar\,1\/} and {\it Bar\,2\/}. \end{enumerate} \section{The model\label{sec:model}} For our model, we assume that the mass distribution in the inner few pc of the Galaxy can be reasonably well approximated by a spherically symmetric ansatz: \begin{equation} \frac{M(R)}{{\rm M}_\odot} = \frac{M_0}{{\rm M}_\odot} + 1.6\,10^{6} \left(\frac{R}{\rm pc}\right)^{1.25} \end{equation} with $R$ the radius from the center of the Galaxy, $M(R)$ the radial mass distribution, and $M_0 = 3\,10^6\,{\rm M}_\odot$ the mass of the central black hole \cite{GEOE97}; the radially dependent contribution is an interpolation given by \citeasnoun{LDB93} for radii smaller than a few hundred pc. We assume the gaseous material to flow on closed circular orbits in the gravitational potential of this mass distribution. In order to allow for local turbulent motion (which however is assumed not to dominate the flow patterns), an additional random, however on average isotropic velocity component is introduced. This corresponds -- through hydrostatic equilibrium -- to a certain thickness of the different flow patterns. As, at this point we are not heading for a self-consistent hydrodynamic model of the flow, we finally allow for an ad hoc radial accretion velocity of 5\% of the local keplarian azimuthal velocity is assumed. We find that the model is much more sensitive to the absolute values of this radial velocity than of the turbulent velocity. However, given that the accretion velocity has a specified preferred direction (inwards) while the turbulent velocity is isotropic, this is not really a surprise. In Fig.\ \ref{fig:eight}, we show the projected velocity vectors for a disklike flow with a turbulent velocity of 40\%\ of the Keplerian velocity, and a turbulent velocity of 5\% of the Keplerian velocity. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure9.ps}} \caption{A representative velocity distribution for the disc model. The length of the arrows corresponds to the absolute value of the velocity.\label{fig:eight}} \end{figure} Choosing the turbulent velocity, i.e., the thickness of the disk, and the accretion velocity, in terms of an accretion disk model would be equivalent to prescribing the value of the viscosity. In terms of the standard theory of accretion disks \citeaffixed{FKR92}{see, e.g.,}, this corresponds to a viscosity parameter $\alpha \sim 0.3$, i.e., a value well within what one finds for other types of disks. We now model the flow by searching for the minimum number of different planes that are required to cover the observed features in the data cube. This obviously requires us not to think in terms of completely filled accretion disks. What we have in mind is a much more transient phenomenon in the following sense: We assume that clumps of gas -- undergoing some type of viscous interaction and/or tidal stretching -- are falling towards the GC. Locally they are described by a flow with Keplerian plus turbulent plus accretion motion. We find that the smallest number of planes to fit the whole measured data cube is three. One of them coincides with that of the CND. We will call that plane the main plane, or plane (i). In Tab.\ \ref{tab:one} we give the orientations of the three planes. While for a single planes, some uncertainty about the sign of the inclination angle would be left, for the ensemble of three planes the consistency indicates the correct choice. \begin{table} \caption{Orientation of the three planes.} \label{tab:one} \begin{tabular}{lrrc} \hline \noalign{\smallskip} & position & & \\ plane & angle & incli- & closest \\ & (N$\rightarrow$E) & nation & edge \\ \noalign{\smallskip} \hline \noalign{\smallskip} (i) & 28$^\circ$ & 25$^\circ$ & W \\ (ii) & 132$^\circ$ & -15$^\circ$ & SW \\ (iii) & 115$^\circ$ & 20$^\circ$ & NE \\ \noalign{\smallskip} \hline \end{tabular} \end{table} To give an idea about the realtive orientations, in Fig.\ \ref{fig:nine} we show the orientation of planes (i) and (iii). The arrows indicate the senses of rotation. The cross marks the position of SgrA*. Plane (i) is the one that is oriented roughly top-bottom. In plane (i) the disk's velocity is positive for positive declination offsets and negative for negative declination offsets from Sgr A*. In the case of plane (iii) there are negative velocities for positive declination offsets and positive velocities for negative declination offsets. Thus one has a kind of counter-rotation of the material in plane (iii) with respect to that in plane (i). \begin{figure} \resizebox{\hsize}{!}{\includegraphics{drehrich.eps}} \caption{The relative orientation of planes (i) (top-bottom) and (iii). The arrows indicate the sense of rotation. The cross shows the position of SgrA*.\label{fig:nine}} \end{figure} \section{Comparing the model with the data} In order to evaluate the quality of the model fit, we determine the residuals of the subtration of the resulting model data cube from the observed one. At the same time, subtracting individual model planes from the entire observed data cube allows to demonstrate where individual structures are located. As an example, in Fig.\ \ref{fig:ten} we show the data cube after subtraction of plane (i). \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure10.ps}} \caption{Residual of the data cube after subtracting plane (i) from it.\label{fig:ten}} \end{figure} When performing this for all three planes, we find that \begin{enumerate} \item the Western Arc, the Northern Arm and the Bar\,1 are contained in plane (i), whereas the whole Eastern Arm and the Bar\,2 are situated beyond; \item plane (ii) covers the Finger of the Eastern Arm and the Bar\,2; \item the rest of the Eastern Arm is situated in plane (iii). \end{enumerate} In Fig.\ \ref{fig:eleven} we give the residual between observed and model data cube after subtracting all three planes and projecting it onto the sky. The only remaining structures are a small part of the Eastern Arm and features in the immediate vicinity of the IRS\,8 complex in the north. All other structures can be represented well within our model of the flow being mainly confined to three planes. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure11a.ps}} \\ \resizebox{\hsize}{!}{\includegraphics{figure11b.ps}} \caption{Contour representation of the original data (top panel) and the final residual (bottom panel) with the same contour levels.\label{fig:eleven}} \end{figure} This indicates that the Western Arc, the Northern Arm and the Bar\,1 are located in the main plane, i.e., the plane of the CND. What in projection appears a one structure called Eastern Arm seems to consist actually of at least two components, one of which lies in the same plane as the Bar\,2, while the remainder is located in a different plane. \section{The distribution of the material in three spatial dimensions} With the knowledge gained in the previous Section, one can now invert the problem and determine the flow patterns in the three planes, independent of the projected features. In Fig.\ \ref{fig:twelve} do so for plane (i), the main plane which coincides with the plane of the CND. In the Fig., we show a view face-onto plane (i). It shows the Northern Arm at the left of the center, the Bar\,1 at the right and the Western Arc which is bent from the end of the Northern Arm to a point below the center. Through projection an observer on Earth see the distribution shown in Fig.\ \ref{fig:thirteen}: One finds the shapes of the Northern Arm and the Western Arc. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure12.ps}} \caption{Contour plot of the distribution of emitting material in plane (i) seen face-onto that plane. The granular structure is an artefact due to discretization in the model.\label{fig:twelve}} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure13.ps}} \caption{The distribution of Fig. \ref{fig:twelve} projected as an observer on Earth sees it.\label{fig:thirteen}} \end{figure} The same procedures were repeated for planes (ii) and (iii). The results are shown in Figs.\ \ref{fig:fourteen}. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure14a.ps}} \\ \resizebox{\hsize}{!}{\includegraphics{figure14b.ps}} \caption{Contour representation the distribution of emitting material in planes (ii) and (iii). Both plans are shown face-on.\label{fig:fourteen}} \end{figure} Figure \ref{fig:fifteen} shows the combined projection of all three planes for an Earth-based observer providing a remarkably nice fit of the measured \neii\ data. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure15.ps}} \caption{The ionized gas distribution of all three planes shown as an observer on Earth sees it.\label{fig:fifteen}} \end{figure} Going to lower intensity levels, we find the possibility of a tenuous gas component which filling almost entirely plane (i). There are hints that this could also be the case for plane (ii), whereas the material in the third plain seems to form unconnected entities (Figure \ref{fig:sixteen}). \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure16.ps}} \caption{Same representation as in Fig. \ref{fig:fifteen}, however for a lower intensity level.\label{fig:sixteen}} \end{figure} Finally, we give in Fig.\ \ref{fig:seventeen} a view of the same data, however seen from a different angle (155$^\circ$ with respect to the $z$-axis) chosen such that planes (i) and (ii) are seen edge-on. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{figure17.ps}} \caption{The same data as in Figure \ref{fig:sixteen} rotated by an angle of 155$^\circ$ with respect to the $z$-axis. Plane (i) is seen vertically edge-on; plane (ii) is also seen edge-on. For an even clearer visualization of the 3D structure of the data cube, we also show it in an animation (\harvardurl{anim2.gif}). \label{fig:seventeen}} \end{figure} \section{Conclusion\label{sec:concl}} We have re-analyzed a \neii\ emission line distribution in the immediate vicinity of the GC by means of a three dimensional visualization of the data cube. We find that due to projection effects some physically distinct entities mislead us by appearing as single features. Based on the data cube, we have re-classified the structures in Sgr A West. We find that they are confined to mainly three distinct planes. Within the planes they represent the denser material. However, at least in two of the three planes, tenuous gas seems to fill almost the entire planes, not unlike accretion disks. Most of the minispiral's material is located in a main plane coinciding with that of the CND. We find the best fit for material moving in the planes on Keplerian circular orbits, overlayed with a turbulent velocity component of some 40\% of the Keplerian speed and an inwards radial velocity of 5\% of the orital velocity. It is important to note that our fits exclude predominant outwards motion; it seems that a scenario of a mass stream from outside towards a central black hole is supported by our results. It is tempting to speculate that the different components that project as the minispiral onto the sky actually are chunks of matter ''falling out of the CND towards the black hole" (see also \citeasnoun{ZG99}). Then the presence of individual planes would be of little surprise. They are -- more or less randomly -- defined by the local angular momentum of the material leaving the CND. One expects -- on average -- a direction coinciding with that of the CND. However, as individual clumps may very well have an angular momentum that differs somewhat from the local average, for individual streams, planes different from that of the CND are possible. At the same time this tells us that these features are rather short-lived. The three dimensional spatial reconstruction of the distribution of the matter moreover indicates that counterrotating features are present. At this point it is suggestive but not yet clear whether this has anything to do with the counter-rotating early type stars found by \citeasnoun{GTKal96}. One can -- as a speculation -- not even exclude that the material in plane (iii) is indeed the rest of the material out of which those stars were formed. \section*{Acknowledgements} We thank J.H.\ Lacy for making his data available to us, and D.A.\ Roberts and W.M.\ Goss for allowing is to use their radio image, in this paper Figure 2. We benefitted much from discussion of the topic of this paper with P.G.\ Mezger. The help of M.-F.\ Landr\'ea in the preparation of the animations is highly acknowledged.
\section{Introduction} Wormholes are nontrivial topological configurations of spacetime that can be represented by solutions of Einstein field equations with stress-energy tensor fields that somewhere violate the so-called average null energy condition (see Ref. \cite{visser-book} for a detailed discussion). Although microscopic violations of the energy conditions are well known (e.g. the Casimir effect), it is far from clear whether stable, macroscopic wormholes can naturally exist in the Universe. One of the ways in which one may obtain violations to the energy conditions is via a scalar fields coupled to gravity (see for instance \cite{NOS1} and references therein). Wormhole formation at a late cosmic time requires Lorentzian topology change in space, something that appears to be more than problematic to most physicists because it implies causality violations \cite{geroch,hawking}. However, if wormholes are created altogether with spacetime and not formed by astrophysical processes, one could expect a cosmological population of these objects without the uncomfortable predictions of topology change theorems. In a couple of recent papers we have discussed the observable effects that could arise from an intergalactic population of natural wormholes \cite{nosotros1,nosotros2}. Since wormhole's mouths could have a total negative mass, they should exert a repulsive gravitational force that can provide very peculiar microlensing events when acting upon the light of compact, background sources \cite{cramer}. Extragalactic wormholes with absolute masses of $\sim$ 1 M$_{\odot}$ would produce very compact Einstein rings, in such a way that just small, ultraluminous sources like the $\gamma$-ray emitting core of quasars (typical size $10^{14}-10^{15}$ cm) might result gravitationally magnified. We have shown in Ref. \cite{nosotros2} that the lightcurve signature of wormhole microlensing events of this sort very much resembles some kinds of gamma ray bursts (GRBs). When a negative mass lens crosses the line of sight to a distant quasar, dragging the caustic pattern along with it, two bursting $\gamma$-ray events will appear in the observer's frame: the first one is the specular image of a fast-rise-exponential-decay (FRED) burst, whereas the second, after a period of stillness that can last several years, is a pure FRED event. In our previous study \cite{nosotros1}, we have used the available database of GRB observations gathered by the BATSE instrument, part of the Compton satellite, to set an upper limit to the amount of negative mass (under the form of compact objects of astrophysical size) in the Universe. Such limit results as low as $|\rho|\leq 2\times10^{-33}$ g cm$^{-3}$ with the most optimistic assumptions. In the present paper we give a step further and embark on the first detailed search for individual wormhole signatures in astronomical databases. GRBs produced by natural wormholes can be differentiated from those originated in fireballs because of two very definite properties: 1) they repeat, and 2) one of the repeating bursts has an anti-FRED time profile, something that cannot be the result of an explosive event \cite{romero} (the companion burst must display a FRED-like lightcurve). We have quantitatively analysed a subsample of the GRBs included in the BATSE 3B cataloge with the aim of identify events that could be unequivocally attributed to wormhole lensing. In what follows we present the results we have obtained. \section{Data analysis} We have analysed a sample of 631 bursts from BATSE 3B catalog whose global symmetry properties were already discussed by Link \& Epstein \cite{link}, and Romero et al. \cite{romero}. This sample contains both faint and bright bursts, spanning 200-fold range in peak flux. PREB + DISC data tapes at 64 ms time resolution, with four energy channels, were used in the analysis. Since the variety of burst profiles is huge and simple visual inspection can be misleading, we have used the skewness function $\cal{A}$ introduced by Link \& Epstein \cite{link} in order to separate those GRBs with anti-FRED profiles. The skewness is basically defined as the third moment of the individual burst time profile and can be directly computed from the observational data as in Ref. \cite{link}. Negative values of $\cal{A}$ correspond to events with slower rising than decaying timescales, thus showing a peculiar asymmetric burst (PAB). In a first step, we estimated $\cal{A}$ for all GRBs in the sample at different background cutoff levels. Just 91 out of 631 bursts present $\cal{A}<$0 at any background. As discussed in Ref. \cite{romero}, most of these events can be explained within the standard fireball model of GRBs \cite{rees}. Just anti-FRED--like single peaked events remain inconsistent with the explosive hypothesis. There are 26 of these GRBs in our sample (4.1 \%). Since wormhole lensing not just provides bursts with $\cal{A}<$0 but also a repetition with specular signature, we have searched for time-space clustering in the sample. We have found that 15 out of 26 candidates (about 60 \%) present companions within error boxes at less than $4^{o}$ (the average positional uncertainty in BATSE catalog). We have estimated the statistical significance of this level of positional coincidences through numerical simulations of random sets of 26 events against a background distribution of 605 GRBs. After 1500 simulations we established that the chance associations expected in the subsample are $13.3\pm2.5$, i.e. there is no need to claim for repetition to explain the observed coincidences at error boxes of $4^o$. However, if positional coincidences separated by less than $1^o$ are considered, we find that 3 out of 26 events present companions. According to a new set of simulations, these results can be attributed to chance only at a $2\sigma$ level. Despite the sample is too scarce to draw any conclusion, it is worth mentioning that when a similar study is carried out with the 91 bursts with $\cal{A}<$0 it is found that there are just 4 positional coincidences at less than $1^o$, at $1\sigma$ confidence level. This could imply that the apparent excess is exclusively associatted to single peaked events, as expected from the microlensing model. In order to detect whether there is some suitable wormhole candidate behind the above mentioned statistical analyses, we turned to the individual single peaked events with $\cal{A}<$0 in a finer search. \section{Results} In Table 1 we list by trigger number all single peaked bursts with $\cal{A}<$0 in our sample. In column 2 we indicate the trigger number of any companion burst in the entire BATSE database within a circle of $4^o$ in radius. Columns 3 and 4 display the temporal and angular distances of pairs of events. A negative value of $\Delta T$ means that the anti-FRED event followed to its companion; such bursts can be eliminated as wormhole candidates, at least over the timescales under consideration here. The final column in the table lists the sign of $\cal{A}$, when defined, for companion bursts that belong to our subsample. Bursts with no entries in this column where not analyzed in the present study, circunscribed to the previously defined set of 631 GRBs. We shall consider as candidates for wormhole microlensing just events with companions that present $\cal{A}>$0 at all levels of background (notice that this is a very restrictive criterion, and eliminates the event mentioned in Ref. \cite{nosotros2}). This left us with only 4 candidates: \#254, \#444, \#1924, and \#2201. A further step can be made now by detecting active galactic nuclei (AGNs) within the error boxes of the bursts. These AGNs would constitute the potential background sources of gamma rays. In Table 2 we list pairs of GRBs along with the AGNs (namely compact QSOs) within the BATSE field. We also indicate the morphological type of the bursts with $\cal{A}>$0. As can be appreciated from this table, three pairs of events present quasars in their fields: \#254, \#444, and \#1924. These are the stronger candidates for wormhole microlensing events in our sample of 631 bursts. None of them, however, can be considered as a certain identification because the profiles of the second bursts in each pair are not exact FREDs despite presenting $\cal{A}>$0 at all levels. These bursts have profiles with some substructure which is not present in the first event of the pair. Although such a fine substructure could be an effect of the different light propagation paths (the light can be exposed during its travel to lensing effects by ordinary matter that might result in a distortion of the original profile \cite{prop}) or even an artifact due to the different orientation of the spacecraft at the detection times, we think that the evidence is not strong enough to claim for an indisputable identification. In the case of the pair \{\#2201, \#2679\}, both bursts are single peaked and present the correct symmetry in their profiles. By other hand, there are no cataloged AGNs in the corresponding sky field. This would not be an insurmountable problem for wormhole microlensing because even very weak and normally undetected QSOs can be enhanced by caustic crossing in such a way as to appear as a bright source during a few seconds \cite{nosotros2}. However, in this particular case, the flux ratio of both bursts (which should be similar as it comes from the same source) is too far from unity as to make a case for the lensing argument. The best candidate in the whole sample is the pair \{\#254, \#2477\}. They present correct symmetry properties and unity flux ratio. The event \#2477 has a substructure that make of it not a perfect FRED, but this, as it was mentioned, could stem from propagation effects. Two AGNs, with redshifts of 0.15 and 1.52, are present within the positional uncertainties.\footnote{It should also be pointed out the FRED--anti-FRED pair, \#688 and \#2788, which is located at the same position in the sky (within 4$^o$) together with 156 QSOs. These two bursts might be produced by two different microlensing phenomena with timescales that span out the BATSE operation time.} The greatest difficulty at present time is the huge positional uncertainty, something that will be significantly improved in ten years time. The results of our search, although not conclusive, are sufficiently suggestive as to encourage new studies over larger samples and with the more accurate detectors of the forthcoming generation of gamma ray satellites. \section{Final comment} Our search for natural wormholes through microlensing was sensitive to timescales up to 3.5 years. Repetition of anti-FRED events over longer scales can not be ruled out. Tegmark et al. \cite{teg} have made a repetition study on the entire BATSE sample concluding that a repetition level of $\sim5$ \% with timescales of a few years is compatible with the current data. Our results show that, if repetition is associated to wormhole microlensing alone, it could reach, at most, a level $\sim 4\%$ over timescales larger than 3.5 years. At shorter scales, wormhole-induced repetitions are constrained at a level $<0.2$ \% (assuming \{\#254, \#2477\} as the sole possible candidate). Since microlensing timescale increases with larger masses of the lenses, the absence of clear detections in our search might be saying that wormholes, if there exist at all, have a mass distribution peaking far beyond the few tenths of solar masses required to produce typical microlensing events with timescales of a few years. If we recall that a negative mass of the size of Jupiter is necessary to keep open a wormhole throat of about 2 m in diameter \cite{visser-book}, this result could be kindly greeted by optimistic interstellar-travel afficionados looking for larger spacetime tunnels. \section*{Acknowledgements} We acknowledge Bennett Link for earlier discussions on burst temporal profiles. The present research has made use of the NASA/IPAC Extragalactic Database, which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. This work has been partially supported by Argentine agencies CONICET (G.E.R., D.F.T.) (also, through research grant PIP 0430/98 and ANPCT, G.E.R.). L.A.A. thanks FOMEC Program for additional support.
\section{Introduction} The work of Einstein, Podolsky, and Rosen (EPR) has started a lasting debate about the completeness and the meaning of local realities in quantum mechanics. In order to make their point, EPR used the following wave function for a system composed of two particles \cite{EPR}: \begin{equation} \label{epr} |{\rm EPR}\rangle = \int \frac{dp}{2\pi} \ |p,-p\rangle = \int dq \ |q,q\rangle. \end{equation} Compared to the original paper of EPR, we have put here spatial separation between the particles equal $q_0=0$. The above wave function describes an entangled state of two particles $a$ and $b$, with the following properties: \begin{equation} (\hat{q}_{a} -\hat{q}_{b})|{\rm EPR}\rangle=0, \ \ \ (\hat{p}_{a} + \hat{p}_{b})|{\rm EPR}\rangle=0. \end{equation} However, further studies of quantum nonlocality and entanglement, especially those providing quantitative tests of compatibility of quantum mechanics with local theories in the form of Bell inequalities, used mainly spin-1/2 particles instead of systems with continuous degrees of freedom. Quantum correlations for position-momentum variables associated with the EPR state (\ref{epr}) can be analyzed in phase space using the Wigner function or the positive $Q$ function. Using the Wigner function approach, Bell \cite{Bell} has argued that the EPR wave function (\ref{epr}) will not exhibit nonlocal effects, because its Wigner function is positive everywhere, and as such will allow for a hidden variable description of the correlations. This statement is correct as long as the measured observables have a straightforward classical interpretation in the phase space representation. However, the situation can change dramatically, if we take into account quantum observables for which the phase space cannot serve as a local theory model. In a recent publication \cite{KBKWEPR} we have demonstrated that the Wigner function of the EPR state, though positive definite, provides a direct evidence of the nonlocal character of this state. The demonstration was based on the fact that the Wigner function can be interpreted as a correlation function for the joint measurement of the parity operator. In this presentation we review the EPR state in the Wigner representation and extend our approach to the positive $Q$ function, which also can be regarded as a correlation function for the joint measurement of certain dichotomic observables. The analysis will be performed for an optical realization of the EPR state as the entangled two-mode state of light generated in the spontaneous parametric process. In this quantum optical context, analysis of the $Q$ function is of particular interest, as the experimental demonstration of nonlocal correlations exhibited by the $Q$ function poses less stringent requirements on single photon detectors used in the setup. This paper is organized as follows. First, in Sec.~2, we briefly review the quantum optical version of the EPR state. In Sec.~3 we discuss the Wigner and the $Q$ functions of this state, and study the limiting case in which the original EPR state is recovered. In Sec.~4 we present the quantum optical scheme which can be used to demonstrate nonlocality of the EPR state in the phase space. Finally, Sec.~5 concludes this presentation. \section{EPR state in quantum optics} Before we show that the phase space representation of the EPR state fully exhibits the nonlocal character of this entangled state, we shall give a brief description of an optical realization of this state, in terms of the two-mode correlated light generated in the spontaneous parametric down-conversion process. With a clear application to quantum optics, it has been shown that a state produced in a process of nondegenerate optical parametric amplification (NOPA) \cite{Reid}, is the optical analog of the EPR state in the limit of strong squeezing. The NOPA process is a nonlinear interaction of two quantized modes (denoted by $a$ and $b$) in a nonlinear medium with a strong classical pump field. The strength of the interaction can be characterized with the parameter $\chi$, which involves the second-order susceptibility and the pump field amplitude. The interaction Hamiltonian of the system is: \begin{equation} H= i \chi (\hat{a}^{\dagger}\hat{b}^{\dagger} - \hat{a}\hat{b}). \end{equation} If the initial state of the system consists of two vacuum modes the NOPA generates: \begin{equation} |{\rm NOPA}\rangle = e^{r(\hat{a}^{\dagger}\hat{b}^{\dagger} - \hat{a}\hat{b})} |0,0\rangle, \end{equation} where $r=\chi t$ is a dimensionless parameter characterizing the interaction time. Simple algebra, based on the following disentanglement of the evolution operator: \begin{equation} e^{r(\hat{a}^{\dagger}\hat{b}^{\dagger} - \hat{a}\hat{b})} = e^{\tanh r \, \hat{a}^{\dagger} \hat{b}^{\dagger}} \left(\frac{1}{\cosh r} \right)^{\hat{a}^{\dagger}\hat{a} +\hat{b}^{\dagger} \hat{b} +1} e^{-\tanh r \, \hat{a}\hat{b}}, \end{equation} shows that the generated state has a diagonal decomposition in terms of the number states of the two modes: \begin{equation} \label{nopa} |{\rm NOPA}\rangle = \frac{1}{\cosh r} \sum_{n=0}^{\infty} (\tanh r)^{n} |n,n\rangle. \end{equation} In order to see the relation between this state and the EPR state (\ref{epr}), we rewrite the state (\ref{nopa}) in the following form: \begin{equation} |{\rm NOPA}\rangle = \frac{1}{\cosh r} \sum_{n=0}^{\infty} (\tanh r)^{n} \int\,dq \int\,dq' \,|q,q'\rangle\langle q,q'| n,n\rangle. \end{equation} Using the fact that the scalar products can be expressed in terms of the Hermite polynomials (in dimensionless units): $\langle q|n\rangle\ = (2^{n}n!\sqrt{\pi})^{-1/2} H_{n}(q)\exp(-q^{2}/2)$, and the following summation formula (valid for arbitrary $\lambda \le 1$): \begin{equation} \label{suma} \sum_{n=0}^{\infty} {\lambda^{n}}\langle q|n\rangle\ \langle n|q'\rangle\; =\;\frac{1}{\sqrt{\pi(1-\lambda^{2}})}\; \exp\left(-\frac{ q^{2}+ q'^{2} -2\lambda qq'}{2(1-\lambda^{2})}\right), \end {equation} we obtain: \begin{equation} \label{nopar} |{\rm NOPA}\rangle = \frac{1}{\sqrt{\pi}} \int\,dq \int\,dq' \exp\left(-\frac{ q^{2}+ q'^{2} -2qq'\tanh r } {2(1-\tanh^2 r)}\right) |q,q'\rangle. \end{equation} This formula is a regularized version of the EPR state (\ref{epr}), with a gaussian smoothing profile of the plane waves. Now it becomes clear, that in the limit of $r \rightarrow \infty$ i.e., for a very long interaction time, this smoothing function becomes a sharp function: $\delta (q-q')$. This follows from the fact that for $\lambda=1$, the sum (\ref{suma}) reduces to the completeness relation for the oscillator wave functions. In this limit the state (\ref{nopar}) becomes exactly the EPR state (\ref{epr}), and as result we obtain, that in terms of photons, the EPR state is: \begin{equation} |{\rm EPR}\rangle \sim \lim_{r\rightarrow \infty} |{\rm NOPA}\rangle \sim |0,0\rangle + |1,1\rangle +|2,2\rangle + \dots \,. \end{equation} The NOPA state has been generated experimentally \cite{Ou} and applied to discuss the implications of the positivity of the corresponding Wigner function on the Bell inequality \cite{OuAPB}. \section{EPR state in phase space} In this section, we discuss the Wigner and the $Q$ functions of the state produced in the nondegenerate spontaneous parametric down-conversion process. We shall pay particular attention to the limit of strong interaction for $r\rightarrow\infty$, where the original, singular EPR state is recovered. \subsection{The Wigner function in the coherent state representation} The two mode Wigner function of the NOPA state (\ref{nopar}) can be calculated directly from the definition: \begin{equation} W(\alpha;\beta) = \int \frac{ d^{2}\alpha'}{\pi^{2}} \int \frac{ d^{2}\beta'}{\pi^{2}} \exp(\alpha \alpha'^{\ast}-\alpha ^{\ast} \alpha' + \beta \beta'^{\ast}-\beta ^{\ast} \beta') \langle \hat{D}_a(\alpha') \hat{D}_b(\beta')\rangle\,, \end{equation} where $\hat{D}_a(\alpha)$ and $\hat{D}_b(\beta)$ are the Glauber's displacement operators for modes $a$ and $b$. Simple calculation shows that the Wigner function of the state (\ref{nopar}) has the form: \begin{equation} \label{nopawig} W(\alpha ;\beta) = \frac{4}{\pi^2} \exp[-2 \cosh 2r (|\alpha|^2 + |\beta|^2) + 2 \sinh 2r (\alpha \beta + \alpha^\ast \beta^\ast)]. \end{equation} This positive everywhere Wigner function is plotted in Fig.~1 for real values of $\alpha$ and $\beta$ and for $r=1$. The Wigner function of the original EPR state (\ref{epr}) is obtained in the limit $r\rightarrow \infty$: \begin{equation} W(\alpha; \beta) \sim \delta(\alpha_{r} -\beta_{r}) \delta(\alpha_{i} +\beta_{i}), \end{equation} where $\alpha_{r}\ (\alpha_{i})$ and $\beta_{r}\ (\beta_{i})$ are the real (imaginary) parts of $\alpha$ and $\beta$. Note the singular form of this Wigner function, due to the singular character of the original EPR wave function (\ref{epr}). \begin{figure} \begin{center} \epsfig{file=wab.ps,bbllx=115,bblly=78,bburx=545,bbury=690,angle=270,% width=12cm} \end{center} \medskip \caption{Fig.~1. Plot of the Wigner function of the regularized EPR state with $r=1$ for real values of $\alpha$ and $\beta$.} \end{figure} \subsection{The $Q$ function in the coherent state representation} The two mode $Q$ function of the NOPA state can be calculated directly from its definition: \begin{equation} Q(\alpha; \beta)= \frac{1}{\pi^{2}} |\langle\alpha,\beta|{\rm NOPA}\rangle|^{2}. \end{equation} It is easy to show that for the NOPA state (\ref{nopar}) this function has the following form: \begin{equation} \label{nopaQ} Q(\alpha;\beta) = \frac{1}{(\pi \cosh r)^{2}} \exp\left(-|\alpha|^{2} -|\beta|^{2} + \tanh r (\alpha^{*} \beta^{*} +\alpha \beta)\right), \end{equation} with the following marginals: \begin{eqnarray} Q(\alpha) = \frac{1}{\pi (\cosh r)^{2}} \exp\left(-\frac{|\alpha|^{2}} {(\cosh r)^{2}} \right)\nonumber\\ Q(\beta) = \frac{1}{\pi (\cosh r)^{2}} \exp\left(-\frac{|\beta|^{2}} {(\cosh r)^{2}}\right). \end{eqnarray} This two-mode $Q$ function is plotted in Fig.~2 for real values of $\alpha$ and $\beta$ and for $r=1$. The $Q$ function of the original EPR state (\ref{epr}) obtained in the limit $r\rightarrow \infty$ is proportional to: \begin{equation} \label{eprQ} Q(\alpha; \beta) \sim \exp(-|\alpha|^{2} - |\beta|^{2} +\alpha^{*} \beta^{*} +\alpha \beta). \end{equation} The last term in the exponent is responsible for the correlation of the entangled EPR state. Note that if we consider the $Q$ function given by Eq.~(\ref{eprQ}) as a limit of Eq.~(\ref{nopaQ}), there is a vanishing prefactor with the magnitude proportional to $e^{-2r}$. The function (\ref{eprQ}) can be derived directly from the EPR wave function (\ref{epr}): \begin{equation} Q(\alpha, \beta)\sim \left|\int dq \langle q,q|\alpha,\beta\rangle \right|^{2}. \end{equation} This simple integration of the coherent state wave functions, in position representation, can be calculated and as a result we reproduce the result (\ref{eprQ}). \begin{figure} \begin{center} \epsfig{file=qab.ps,bbllx=115,bblly=78,bburx=545,bbury=690,angle=270,% width=12cm} \end{center} \medskip \caption{Fig.~2. Plot of the $Q$ function of the regularized EPR state for real values of $\alpha$ and $\beta$, for $r=1$.} \end{figure} \section{Testing quantum nonlocality in phase space} In this presentation we propose to probe the correlations of the NOPA state with a scheme involving two photon counting detectors for modes $a$ and $b$. The setup to demonstrate quantum nonlocality for the Wigner function and the $Q$ function is presented in Fig.~3. The photons from the NOPA source are measured by two photon counting detectors preceded by two beam splitters. These beam splitters, with the transmission coefficients very close to one, mix the NOPA photons with two highly excited coherent states. It has been shown elsewhere \cite{KBKWPRL96}, that in the limit of the transmission tending to one, the effect of the beam splitters is equivalent to phase space shifts for the two modes by $\hat{D}_a(\alpha)$ and $\hat{D}_b(\beta)$, where $\alpha$ and $\beta$ are complex amplitudes of the reflected coherent fields. A single-mode version of such an experiment, involving measurements of the displaced photon number statistics for simple classical states of light, is reported in these proceedings in the context of quantum state measurement \cite{ExpWig}. We shall show that if the detectors placed in the two arms of the proposed measurement scheme can resolve the number of absorbed photons, the phase space Wigner function can be determined directly. Moreover, it will become apparent that the Wigner function measured in this setup describes correlations between the parity of the number of photons registered by the two detectors. As the parity is a dichotomic $\pm 1$ variable, the Wigner function can be therefore directly inserted into appropriate Bell inequalities in order to test the nonlocal character of the NOPA state. The most efficient single-photon detectors available currently are avalanche photodiodes operated in the Geiger regime. These detectors cannot resolve the number of simultaneously absorbed photons, and provide only a binary answer saying whether any photons have been registered or not \cite{Kwiat}. We shall show that use of these detectors in our scheme leads in a natural way to the measurement of the two-mode $Q$ function, and that the binary answer of the detectors makes the $Q$ function a nonlocal correlation function which violates the corresponding Bell inequality. \begin{figure} \begin{center} \epsfig{file=nopa.ps,bbllx=130,bblly=340,bburx=460,bbury=540,width=10cm} \end{center} \medskip \caption{Fig.~3. The experimental scheme proposed to test nonlocality exhibited by the quantum optical realisation of the EPR state. The nonlinear crystal NL pumped by the coherent field generates two correlated beams. The high-transmission beam splitters with auxiliary coherent fields placed in the paths of these beams perform the displacement transformations $\hat{D}_a(\alpha)$ and $\hat{D}_b(\beta)$. The phase space displaced beams are monitored by the photon counting detectors PC1 and PC2.} \end{figure} \subsection{Violation of local reality by the Wigner function} In order to relate the Wigner function to photocount correlations, we will consider the case when the detectors are capable of resolving the number of absorbed photons. Let us assign to each event $+1$ or $-1$, depending on whether an even or an odd number of photons has been registered. In this case the joint correlation for the photon number parities is: \begin{equation} E(\alpha ;\beta) = \langle{\rm NOPA}| \hat{\Pi}_{a}(\alpha) \otimes \hat{\Pi}_{b}(\beta) |{\rm NOPA}\rangle, \end{equation} where \begin{equation} \hat{\Pi}_{a}(\alpha) = \hat{D}_a(\alpha) (-1)^{\hat{n}_a} \hat{D}^{\dagger}_a(\alpha), \ \ \ \ \ \hat{\Pi}_{b}(\beta) = \hat{D}_b(\beta) (-1)^{\hat{n}_b} \hat{D}^{\dagger}_b(\beta), \end{equation} are parity operators of the modes $a$ and $b$ characterized by the corresponding photon number operators $\hat{n}_a $ and $\hat{n}_b $. Equivalently, $E(\alpha ;\beta)$ can be written as the overall parity operator displaced in the two-mode Hilbert space by the operation $\hat{D}^{\dagger}_a(\alpha) \otimes \hat{D}^{\dagger}_b(\beta)$. This observable is known in the phase space representation to provide, up to the normalization constant, the Wigner function of the quantum state \cite{KBKWPRL96}. On the other hand, we can establish a clear analogy of this scheme with joint spin-1/2 measurements. The counterpart of the polarizer orientation is now the coherent displacement, which can be set freely in each of the two spatially separated apparatuses. Furthermore, classification of the registered number of photons according to the parity provides a dichotomic $\pm 1$ outcome which is analogous to the spin direction. As a result all types of Bell's inequalities derived for spin systems can be used to test the nonlocality of the NOPA Wigner function. In Ref.~\cite{KBKWEPR} we have shown that the Wigner function (\ref{nopawig}) leads to a violation of the the Bell inequality. Here we will just quote the result. For local hidden variable theories the following combination: \begin{equation} {\cal B} = E(\alpha';\beta') + E(\alpha';\beta)+E(\alpha;\beta')- E(\alpha;\beta) \end{equation} should satisfy the Bell inequality $-2\le {\cal B} \le 2$. We have found that for a certain selection of coherent displacements the value of this combination in the limit $r\rightarrow \infty$ is: ${\cal B} = 1 + 3 \cdot 2^{-4/3} \approx 2.19$, which clearly violates the Bell inequality. \subsection{Violation of local reality by the $Q$ function} Based on the previous result we shall demonstrate how nonlocality of the EPR state can be revealed using the positive definite $Q$ quasidistribution function. The quantum optical experiment we shall propose does not require detectors that can resolve the number of photons triggering the output signal. We will be interested in events when {\em no photons} were registered. This measurement is described by the following observable: \begin{equation} \hat{O}_{a}=\lim_{\epsilon\rightarrow 0} \left( \epsilon\right)^{\hat{n}_{a}} = \; : \! e^{-\hat{n}_{a}} \! : \; = |0_{a}\rangle\langle 0_{a}|, \end{equation} for the mode $a$, and a similar observable $\hat{O}_{b}$ for the mode $b$. As we discussed in the previous subsection, these observables can be shifted in phase space using auxiliary coherent fields and high-transmission beam splitters. As a result of this shifting we obtain: \begin{equation} \hat{O}_{a}(\alpha) = \hat{D}_a(\alpha) \hat{O}_{a} \hat{D}^{\dagger}_a(\alpha), \ \ \ \hat{O}_{b}(\beta) = \hat{D}_b(\beta) \hat{O}_{b} \hat{D}^{\dagger}_b(\beta). \end{equation} Obviously, these observables describe projections on a coherent state $|\alpha\rangle\langle\alpha|$ for the mode $a$ and $|\beta\rangle\langle\beta|$ for the mode $b$. Therefore, statistics of no-count events yields directly the $Q$ function of the measured field. \begin{figure} \begin{center} \epsfig{file=viol.ps,bbllx=115,bblly=78,bburx=545,bbury=690,angle=270,% width=12cm} \end{center} \medskip \caption{Fig.~4. Plot of the combination ${\cal CH}$ as a function of the displacement intensity $J$ and the interaction parameter $r$. Only values exceeding the upper bound imposed on ${\cal CH}$ by local theories are shown.} \end{figure} The quantum mechanical probability of no-count events in both the detectors is: \begin{eqnarray} p_{ab}(\alpha;\beta) & = & \langle{\rm NOPA}| \hat{O}_{a}(\alpha) \otimes \hat{O}_{b}(\beta) |{\rm NOPA}\rangle \nonumber \\ & =& \frac{1}{ (\cosh r)^{2}} \exp\left( -|\alpha|^{2} -|\beta|^{2} + \tanh r (\alpha^{*} \beta^{*} +\alpha \beta)\right) \end{eqnarray} where $\alpha$ and $\beta$ are the coherent displacements for the modes $a$ and $b$, respectively. The probabilities of no-count events on single detectors are: \begin{eqnarray} p_{a}(\alpha) & = & \langle {\rm NOPA} | \hat{Q}_{a}(\alpha) \otimes \hat{I}_{b} |{\rm NOPA} \rangle = \frac{1}{ (\cosh r)^{2}} \exp\left(-\frac{|\alpha|^{2}} {(\cosh r)^{2}}\right), \nonumber \\ p_{b}(\beta) & = & \langle {\rm NOPA} | \hat{I}_{a} \otimes \hat{O}_{b}(\beta) | {\rm NOPA}\rangle = \frac{1}{ (\cosh r)^{2}} \exp\left(-\frac{|\beta|^{2}} {(\cosh r)^{2}}\right). \end{eqnarray} We recognize that these probabilities are equal up to a normalization factor the two-mode $Q$ function (\ref{nopaQ}) and its two marginals for the NOPA state. The correlation function $p_{ab}(\alpha,\beta)$ describes a binary $0$ or $1$ measurement on the modes $a$ and $b$ with adjustable parameters of the apparatuses represented by $\alpha$ and $\beta$. As a result of this, the Bell inequality derived for a measurement of local realities bounded by $0$ and $1$, can be applied to test the nonlocal character of the NOPA state (\ref{nopa}), using the $Q$ function. We shall consider the the Clauser-Horne combination \cite{CH} for a selected set of four displacements: \begin{equation} {\cal CH} = p_{ab}(0;0)+ p_{ab}(\alpha;0) + p_{ab}(0;\beta) - p_{ab}(\alpha;\beta)- p_{a}(0) - p_{b}(0), \end{equation} which for local theories satisfies the inequality $-1 \le {\cal CH} \le 0$. We will take the coherent displacements to be real with $\alpha = -\beta =\sqrt{J}$. For these values we obtain \begin{equation} {\cal CH} = \frac{1}{ (\cosh r)^{2}}\left(2 e^{-J} -e^{-2 J(1+ \tanh r)} -1\right). \end{equation} As depicted in Fig.~4, this result violates the upper bound imposed by local theories. In the limit of $r \rightarrow \infty$ and small values of the displacement, this function becomes: \begin{equation} {\cal CH}\approx 8J e^{-2r}, \end{equation} i.e., it violates the Bell inequality for all values of $r$, but the violation becomes infinitesimally small. This results from the fact that the $Q$ function (\ref{eprQ}) in this limit is dumped by the prefactor $1/\cosh^2 r$, decreasing like $e^{-2r}$. \section{Conclusions} In this presentation we have shown that nonlocality of the EPR state can be revealed using its phase space representation of the form of the Wigner or $Q$ quasidistribution functions. This is possible owing to the observation that these quasidistribution functions describe nonlocal correlations that can be detected in a certain quantum optical scheme. Of course, this scheme is much more general. It can be applied to measure an arbitrary two-mode state of light, and the corresponding Wigner and $Q$ functions will always have the operational meaning of nonlocal correlation functions \cite{KBKWPRL99}. \section*{Acknowledgements} This work has been partially supported by the Polish KBN grant No. 2 P03B 089 16. K.W. thanks the Alexander von Humboldt Foundation for generous support and Prof.\ W.~P.~Schleich for his hospitality in Ulm.
\section{Introduction} The recent atmospheric neutrino data from the Super-Kamiokande (SK) experiment \cite{To98} can be beautifully explained through flavor oscillations generated by nonzero neutrino mass and mixing \cite{Po67,Ka62} in the $\nu_\mu\leftrightarrow\nu_\tau$ channel \cite{SKEV}. Such interpretation is consistent with all the SK data, including sub-GeV $e$-like and $\mu$-like events (SG$e,\mu$) \cite{SKSG}, multi-GeV $e$-like and $\mu$-like events (MG$e,\mu$) \cite{SKMG}, and upward-going muon events (UP$\mu$) \cite{SKUP}. A combined analysis of the 33 kTy SK data sample can be found in \cite{Fo99}. The oscillation hypothesis has been strengthened by the latest (preliminary) 45 kTy SK data sample \cite{Me99,Ha99}, and is also consistent with independent atmospheric neutrino results from the MACRO \cite{MACR} and Soudan-2 \cite{SOUD} experiments, as well as with the finalized upward-going muon data from the pioneering Kamiokande experiment \cite{KAUP}. Establishing $\nu_\mu\leftrightarrow\nu_\tau$ oscillations generated by nonzero $\nu$ mass and mixing as the ``standard'' interpretation requires, however, further data and analyses. Basically, the following three aspects should be clarified: (1) the periodicity; (2) the flavors; and (3) the dynamics. So far, the periodicity of the $\nu_\mu$ oscillation pattern has not been experimentally observed in the neutrino energy $(E)$ or pathlength $(L)$ domain, and it is unlikely to emerge from the SK lepton distributions, largely smeared in energy or angle. Although specific nonperiodic scenarios, such as neutrino decay \cite{Ba99}, can be indirectly excluded by careful analyses of SK data \cite{Li99,Sc99}, the direct observation of a periodic disappearance pattern of $\nu_\mu$'s remains an important goal for future atmospheric and long-baseline neutrino experiments. Besides periodicity, one should also identify unambiguously the flavor(s) of the oscillating partner(s) of $\nu_\mu$'s because, in principle, all oscillation channels into active or sterile neutrinos might be open ($\nu_\mu\leftrightarrow\nu_\tau$, $\nu_\mu\leftrightarrow\nu_e$, $\nu_\mu\leftrightarrow\nu_s$). While the amplitude of possible $\nu_\mu\leftrightarrow\nu_e$ transitions is bound to be small \cite{Fo99}, one cannot exclude $\nu_\mu\leftrightarrow\nu_s$ oscillations with large amplitude with present SK data \cite{Me99,Ha99}. However, based on the fact that different oscillation channels induce somewhat different energy-angle lepton distributions, there are good prospects for a significant ($>3\sigma$) discrimination of $\nu_\mu\leftrightarrow\nu_\tau$ from $\nu_\mu\leftrightarrow\nu_s$ with future SK data \cite{Ka99}. In this work we assume that both the periodicity (i.e., the existence of an oscillation length $\lambda$) and the oscillating flavors ($\nu_\mu,\nu_\tau$) are established, and we rather focus on the third aspect to be clarified: The dynamical origin of $\nu_\mu\leftrightarrow\nu_\tau$ oscillations. The standard oscillation dynamics, involving a nontrivial $2\times2$ neutrino mass matrix, leads to a well-known energy dependence of $\lambda$, \begin{equation} \lambda^{-1}\propto E^{-1}\ {\rm (standard)}\ . \label{std} \end{equation} However, possible nonstandard neutrino interactions or properties can also generate (or coexist with) $\nu_\mu\leftrightarrow\nu_\tau$ oscillations \cite{Gr95,Jo98}. An incomplete list of possibilities include violations of the equivalence principle (VEP) \cite{Ga88,Ga89}, flavor-changing neutral currents (FCNC) (see \cite{Go99} and refs.\ therein), neutrino couplings to space-time torsion fields \cite{DS81}, neutrino interactions through charged scalar particles \cite{Fu88}, nonrelativistic heavy neutrinos \cite{Ah97}, violations of Lorentz invariance (VLI) \cite{Co97,Gl97} and of CPT symmetry \cite{Co98}. In several such models the energy dependence of $\lambda$ takes a power-law form \cite{Ya94}, \begin{equation} \lambda^{-1}\propto E^{n}\ (n\neq -1,{\rm\ nonstandard)}\ . \label{nonstd} \end{equation} Although models with exotic dynamics for $\nu_\mu\leftrightarrow\nu_\tau$ oscillations do not survive Occam's razor, they might survive experimental tests! Effective tests must cover the widest possible energy range, as evident from Eqs.~(\ref{std}) and (\ref{nonstd}). Concerning atmospheric neutrinos, pre-SK data analyses covered only about two decades in energy (i.e., the so-called contained events, $E\sim 0.1$--$10$ GeV), and were compatible with several nonstandard scenarios, in particular with the VEP hypothesis (corresponding to $n=1$) \cite{Pa93,Ha96}. An interesting post-SK analysis, covering a slightly more extended energy range \cite{Fo98}, does not appear to discriminate significantly the three cases $n=0$ and $n=\pm1$ examined. However, as observed in \cite{Li99,Sc99}, a much longer ``lever arm'' in the energy domain is provided by the inclusion of partially contained and upward-going muon events (up to $E\sim 10^3$ GeV), thus providing a powerful tool to test exotic scenarios (which, indeed, appear to be disfavored in general \cite{Li99}). In this work we assess quantitatively the situation for $\nu_\mu\leftrightarrow\nu_\tau$ models with power-law energy dependence of the oscillation length ($\lambda^{-1}\propto E^n$), including, as relevant subcases, standard mass-mixing oscillations $(n=-1)$ and violations of the equivalence principle or of Lorentz invariance $(n=1)$. We obtain two basic results: (1) The 90\% C.L. range of $n$ is determined to be $n=-0.9\pm0.4$, which is in striking agreement with standard oscillations, and excludes all $n\neq -1$ models as dominant sources of $\nu_\mu\leftrightarrow\nu_\tau$ transitions; (2) assuming the $n=1$ case as a subleading mechanism (coexisting with leading standard oscillations), we place stringent upper bounds on its amplitude. Such bounds can be interpreted as upper limits to violations of special or general relativity principles. The plan of the paper is as follows. In Sec.~II we analyze models with generic $\lambda^{-1}\propto E^n$ behavior, and constrain $n$ through fits to SK atmospheric $\nu$ data. In Sec.~III we consider a more complicated case with two coexisting sources for oscillations, namely, neutrino masses and violations of relativity principles. We summarize our results in Sec.~IV. \section{Analysis of models with $\lambda^{-1}\propto E^{\lowercase{n}}$} In this section, we first review some oscillation models with $\lambda$ depending on $E$ through a power law. Then, independently of specific models, we study the phenomenology of atmospheric $\nu$'s, and constrain the energy exponent $n$ through detailed fits to the SK data. \subsection{Review of models} Typical two-flavor hamiltonians for $\nu_\mu\leftrightarrow\nu_\tau$ oscillations predict a flavor transition probability of the form \begin{equation} P(\nu_\mu\leftrightarrow\nu_\tau)= \sin^2 2\xi \, \sin^2 (\pi \lambda^{-1}L)\ , \label{P} \end{equation} where $\xi$ is the rotation (mixing) angle between the flavor basis and the basis where the hamiltonian is diagonal, $L$ is the neutrino pathlength, and $\lambda$ is the neutrino oscillation length. In several cases of interest, $\lambda$ depends on the neutrino energy $E$ as \cite{Ya94,Fo98} \begin{equation} \lambda^{-1} \propto E^n \label{lambda} \end{equation} with the exponent $n$ taking integer (positive or negative) values. For instance, the cases $n=-1$, 0 and 1 arise, respectively, in the presence of neutrino interactions mediated by scalar, vector, and tensor fields \cite{Fo98}. Specific models include: \begin{mathletters} \begin{eqnarray} \lambda^{-1}&=&\frac{\Delta m^2}{4\pi E}\ \ \ \ \ \ \ \ \ (n=-1, {\rm\ standard\ oscillations\ }\protect\cite{Po67})\ , \label{a}\\ &=& \frac{E|\phi|\Delta\gamma}{\pi} \ \ \ \ \ (n=1, {\rm\ violations\ of\ equivalence\ principle\ }\protect\cite{Ga88})\ , \label{b}\\ &=& \frac{E \delta v}{2\pi} \ \ \ \ \ \ \ \ \ \ (n=1, {\rm\ violations\ of\ Lorentz\ invariance\ }\protect\cite{Co97})\ , \label{c}\\ &=& \frac{\delta b}{2\pi} \ \ \ \ \ \ \ \ \ \ \ \ (n=0, {\rm\ violations\ of\ CPT\ symmetry\ }\protect\cite{Co98})\ , \label{d}\\ &=& \frac{Q\delta k}{2\pi} \ \ \ \ \ \ \ \ \ \ (n=0, {\rm\ nonuniversal\ coupling\ to\ a\ torsion\ field\ }\protect\cite{DS81})\ . \label{e} \end{eqnarray} \end{mathletters} Standard oscillations [Eq.~(\ref{a})] are simply generated by nonzero neutrino masses ($\Delta m^2=m^2_2-m^2_1$), with a mixing angle $\xi$ [Eq.~(\ref{P})] usually denoted as $\theta_{23}$, $\psi$, or simply $\theta$ (we adopt the symbol $\psi$ in the following, as we use $\theta$ for the lepton zenith angle). In this case, the energy exponent is $n=-1$. Equation (\ref{b}) refers to possible violations of the equivalence principle (VEP) \cite{Ga88}, namely, to nonuniversal couplings of neutrinos to the gravitational potential $\phi$. The difference in such couplings is usually denoted by $\Delta\gamma$, and the mixing angle by $\theta_G$. The potential seems to be dominated by the local supercluster ($\phi\sim 3\times 10^{-5}$ \cite{Ke90}), but ambiguities in its definition suggest to use the product $\phi\Delta\gamma$ as a single, dimensionless free parameter, rather than $\phi$ and $\Delta\gamma$ separately (see \cite{Ha96,Ba95,Mu97} for recent discussions). The energy exponent of VEP-induced oscillations is $n=1$.% \footnote{An alternative, string-inspired VEP mechanism \protect\cite{Da94}, leading to an energy exponent $n=-1$ rather than $n=1$, has been recently considered in \protect\cite{Ha98}. Its phenomenology would be indistinguishable from the standard case, as far as $\nu_\mu\leftrightarrow\nu_\tau$ oscillations are involved. We do not consider such VEP scenario in this work.} Equation (\ref{c}) refers to possible violations of Lorentz invariance (VLI), namely, to asymptotic neutrino velocities different from $c$ \cite{Co97}. The parameter $\delta v$ represents the speed difference in units of $c$. The mixing angle is usually denoted as $\theta_v$. Since the energy exponent $n$ is the same ($=1$) in both the VEP and the VLI scenario, such mechanisms are phenomenologically equivalent through the substitutions $|\phi|\Delta\gamma \to \delta v/2$ and $\theta_G\to\theta_v$ \cite{Gl97}. Equation (\ref{d}) refers to possible violations of the CPT symmetry through a more general class of Lorentz-violating perturbations \cite{Co98}, $\delta b$ being proportional to the CPT-odd hamiltonian producing energy independent $(n=0)$ oscillations. It is interesting to notice that the most general Hamiltonian considered in \cite{Co98} encompasses the three scenarios with $n=-1$, 0 and 1. Finally, Eq.~(\ref{e}) refers to possible nonuniversal couplings $(\Delta k\neq 0)$ of neutrinos to a space-time torsion field of strength $Q$ \cite{DS81}, which also produce energy-independent oscillations. In principle, several oscillation mechanisms with different $n$'s might occur at the same time, with corresponding complications in the analysis. In the next two subsections, however, we consider only one mechanism at a time. We will discuss the interesting case of coexisting $n=-1$ plus $n=+1$ oscillations in Sec.~III. A remark is in order. Not all ``exotic'' models for $\nu_\mu\leftrightarrow\nu_\tau$ transitions can be parametrized as in Eqs.~(\ref{P}) and (\ref{lambda}). An important exception is represented by FCNC-induced oscillations (see \cite{Go99} and references therein). In fact, although the FCNC oscillation phase is energy-independent $(n=0)$, it is proportional to the column density of electrons rather than to $L$. Therefore, FCNC-induced oscillations deserve a separate analysis \cite{Go99,Li99} and will be considered in a future work. However, some features of the $n=0$ case also apply qualitatively to the FCNC case. Finally, we note in passing that the analysis of nonstandard energy dependences for $\nu$ oscillations has some correspondence in the neutral kaon system, where the role of $\lambda$ is played by the effective ${\rm K}{}^0$--$\overline{\rm K}{}^0$ mass difference (see \cite{El99} and references therein). \subsection{Model-independent analysis} In this section we do not commit ourselves to any specific model, and rather analyze the phenomenology of the $\lambda^{-1}\propto E^n$ dependence in the most general way. We assume that the $\nu_\mu\leftrightarrow\nu_\tau$ oscillation probability takes the form \begin{equation} P(\nu_\mu\leftrightarrow\nu_\tau)=\alpha \cdot \sin^2 \left( \beta\;\frac{L}{10^3{\rm\ km}}\;\frac{E^n}{{\rm GeV}^n}\right)\ , \label{psk} \end{equation} where $\alpha$ is an overall scale factor ($0\leq \alpha\leq 1$) for the oscillation amplitude, $\beta$ is an unconstrained scale factor for the oscillation phase, and $n$ is a free exponent (not necessarily equal to $-1$, 0, or 1). The units in Eq.~(\ref{psk}) have been conveniently chosen on the basis of the current Super-Kamiokande data, which suggest an oscillation length of $O(10^3{\rm\ km})$ for neutrino energies of $O(1 {\rm\ GeV})$. The standard oscillation case is recovered by taking $n=-1$, $\alpha=\sin^2 2\psi$, and $\beta =1.27\, \Delta m^2/(10^{-3} {\rm\ eV}^2)$. Equation~(\ref{psk}) is used to calculate the observable rates of sub-GeV, multi-GeV, and upward through-going muon events in SK as a function of the lepton zenith angle $\theta$, using the same detailed and accurate approach as in Refs.~\cite{Fo99,Sc99}, to which we refer the reader for technical details. Here we just remind that the parent neutrino distributions for sub-GeV (SG), multi-GeV fully contained (MG~FC), multi-GeV partially contained (MG~PC), and upward through-going (UP) muon events are roughly distributed in the ranges $0.1$--$3$ GeV, $1$--$10$ GeV, $1$--$10^2$ GeV, and $10$--$10^3$ GeV, respectively, thus covering about four decades in $E$. Figure~1 shows the expected zenith distributions of SG$\mu$, MG$\mu$ (FC+PC), and UP$\mu$ events in Super-Kamiokande for maximal mixing ($\alpha=1$) and for three representative values of $\beta$: dotted line, $\beta=0.2$; dashed line, $\beta=1$; solid line, $\beta=5$ (on color printers such lines appear in green, red, and blue, respectively). In each bin, the muon rates $\mu$ are normalized to the expectations $\mu_0$ in the absence of oscillations, so that deviations from the no-oscillation case ($\mu/\mu_0=1$) are immediately recognizable. Since we are considering pure $\nu_\mu\leftrightarrow\nu_\tau$ oscillations ($e/e_0=1$ always), the electron rates are not shown. The effect of taking $\alpha<1$ can be approximately described by reducing proportionally the amount of muon disappearance in each subfigure. The theoretical expectations are affected by relatively large and strongly correlated uncertainties (not shown), mainly related to the overall normalization of the atmospheric neutrino flux \cite{Fo99}. The 45 kTy SK data \cite{Me99,Ha99} (dots with statistical error bars) are superposed to guide the eye, although no fit to the data is implied by this figure. In Fig.~1, from top to bottom, the energy exponent $n$ takes the values $n=-2$, $-1$, 0, and 1, corresponding to a dependence of the kind $L/E^2$, $L/E$, $L$, and $L\cdot E$ for the oscillation phase. Such four cases are described in the following. {\em Case $L/E^2$.} This case ($n=-2$) does not correspond to any known model, and is used only to extend the study to a power-law energy dependence faster than in the standard case ($n=-1$). The muons appear to be significantly suppressed at the lowest (SG) energies, the more the larger $\beta$. An excessive suppression of SG muons is avoided only by keeping $\beta \lesssim 1$. On the other hand, such values of $\beta$ are too low to produce a significant suppression of MG muons, which rather prefer $\beta \gtrsim 5$. Therefore, one expects a ``compromise'' between underestimated SG$\mu$ rate and overestimated MG$\mu$ rates for $\beta$ in the range $\sim 1$--$5$. UP$\mu$ events are basically unsuppressed, due to their high energy. Reducing $\alpha$ would be of no help in reducing the conflict between SG and MG muon expectations. {\em Case $L/E$.} This is the ``standard'' oscillation case, which, as well known, provides and excellent description of the SK data for $\Delta m^2\sim {\rm\ few}\times 10^{-3}$ eV$^2$ \cite{SKEV,Fo99}. Therefore, it is not surprising to see in Fig.~1 that, for values of $\beta$ in the range $\sim 1$--$5$, all the muon data are well reproduced. Higher values of $\beta$ would suppress SG muons too much, while lower values would not produce enough suppression at higher energies (MG$\mu$ and UP$\mu$ samples). {\em Case $L$.} In this case, the $\nu_\mu$ survival probability is energy independent, and the differences in the muon suppression patterns among the SG, MG and UP muon samples are mainly due to the different angular smearing. At low (SG) energies the smearing is very effective and there is little difference between the three curves, while higher-energy, MG muons are more discriminating. Values of $\beta$ around unity (dashed curve) seem to be preferred by SG+MG data, and produce a significant $(\sim 1/2)$ suppression below the horizon. UP muons have a different suppression pattern, which is highly correlated with the parent neutrino direction, and follows closely the variations of $P_{\mu\mu}$ with $L$ (dashed and dotted curves), until the oscillations are so fast to be unresolved within the bin width (solid curve). The expected UP$\mu$ suppression appears to be larger than suggested by the data, unless $\alpha$ is taken to be nonmaximal; however, for $\alpha<1$ the relatively good description of SG+MG data would be spoiled (not shown). {\em Case $L\cdot E$.} In this case, the $\nu_\mu$ survival probability rapidly approaches the average value $1/2$ as the energy increases. Therefore, although SG+MG data can be described relatively well with $\alpha=1$ and $\beta\sim 0.2$, the expected UP$\mu$ rates are too low in any case. As in the previous $n=0$ scenario, taking $\alpha<1$ would help to reproduce the UP$\mu$ data, but would worsen the description of SG+MG data. No satisfactory compromise can be reached. In all the four cases, it can be seen that SG (or even SG+MG) data alone do not discriminate strongly among the various scenarios, since they do not probe the full energy range explorable by SK. This might explain why the SG+MG analysis in \cite{Fo98} could not significantly distinguish the three cases $n=-1$, 0, and 1. The inclusion of UP$\mu$ data extends the range up to $E\sim 10^3$ GeV, and provides a very important tool to probe the energy dependence of $\lambda$ \cite{Li99}. In conclusion, from the examination and the comparison of the oscillated muon distributions at different values of $n$ shown in Fig.~1, the case $n=-1$ emerges as a good description of the SK data at all energies, while for $n\neq -1$ the patterns of muon suppression at low (SG), intermediate (MG) and high (UP) energies appear to be in conflict with the data. \subsection{Fits to the Super-Kamiokande data} The qualitative understanding of the muon distributions at different values of $n$ (previous subsection) can be improved by performing quantitative fits to the 45 kTy preliminary SK data \cite{Me99,Ha99}. We use a $\chi^2$ approach that, as described in \cite{Fo99}, takes into account several sources of correlation among the systematics affecting the theoretical predictions. Even if $\nu_e$'s do not participate to oscillation in the scenarios considered here, we have included the SG and MG $e$-like data in the analysis (for a total of 30 data points), since they play an important role in constraining the overall normalization uncertainty (see also \cite{Fo99,Sc99}). Figure~2 shows the best-fit zenith distributions of muons for the four cases considered in Fig.~1 ($n=-2$, $-1$, 0, and 1), as obtained by leaving $\alpha$ and $\beta$ unconstrained. The values of $\alpha$, $\beta$, and $\chi^2$ at the best-fit points are reported in the top part of the figure. The $L/E$ ($n=-1$) case provides an excellent fit to the data ($\chi^2_{\rm min}/N_{\rm DF}=20.3/28$), while all the other cases do not provide a good description of at least one data sample (SG$\mu$, MG$\mu$, or UP$\mu$). In particular, for $n=-2$ there is insufficient up-down asymmetry of MG$\mu$'s and no slope of UP$\mu$'s; for $n=0$ none of the zenith distributions is correctly reproduced; for $n=1$ there is a too strong and flat suppression of UP$\mu$'s. In Fig.~3 we present the $\chi^2$ curve as obtained by taking also $n$ as a free parameter, besides $\alpha$ and $\beta$. Although noninteger values of $n$ may not be related to any realistic oscillation dynamics, this exercise is useful to see how accurately $n$ is determined through the SK data. The result is striking: \begin{equation} n=-0.9\pm0.4 {\rm\ \ at\ 90\%\ C.L.} \label{n} \end{equation} (corresponding to $\chi^2-\chi^2_{\rm min }= 6.25$ for $N_{\rm DF}=3$). This narrow range for $n$ is perfectly consistent with standard ($n=-1$) neutrino oscillations, and inconsistent with any other integer value of $n$. Given the importance of standard $\nu_\mu\leftrightarrow\nu_\tau$ oscillations, we show in Fig.~4 the updated limits on the oscillation parameters $\Delta m^2$ and $\sin^2 2\psi$. We find the best fit at $\Delta m^2=2.8\times 10^{-3}$ eV$^2$ and maximal mixing. The bounds in Fig.~4 are in good agreement with the latest full data analysis from the SK collaboration \cite{Ha99}. In conclusion, standard oscillations $(\lambda^{-1}\propto E^{-1})$ are strongly favored as the {\em dominant\/} mechanism for the $\nu_\mu\leftrightarrow\nu_\tau$ flavor transitions of atmospheric neutrinos in SK. Alternative mechanisms of the kind $\lambda^{-1}\propto E^{n}$ (with $n\neq -1)$ {\em cannot\/} be the dominant source of muon disappearance in SK. In particular, violations of special or general relativity principles (VLI or VPE, leading to $\lambda^{-1} \propto E$) cannot explain the bulk of SK atmospheric $\nu$ data. Therefore, if $n\neq -1$ oscillations occur in nature, they can only be subleading processes with small amplitude, coexisting with leading, large-amplitude $n=-1$ standard oscillations. Such results generalize and refine previous indications that SK data could disfavor some exotic models \cite{Li99}. \section{Constraints on violations of relativity principles} In this section we consider a more complicated case, characterized by leading $n=-1$ oscillations plus subleading $n=+1$ oscillations, possibly generated by violations of relativity principles (VEP or VLI). We show that the fit to the SK data is not improved with respect to the case of standard $\nu_\mu\leftrightarrow\nu_\tau$ transitions. As a consequence, we derive upper bounds on such violations. A brief review of the theoretical formalism precedes the phenomenological analysis. \subsection{Violations of relativity principles: Formalisms} The theory and phenomenology of neutrino violations of the equivalence principle (VEP) \cite{Ga88} have been investigated in a number of papers, including studies of the solar $\nu$ deficit \cite{Pa93,Ha96,Ba95,Mu97,Ha91,Bu93,Mi95,Mu96,Ma98,Ca99}, of the atmospheric $\nu$ anomaly \cite{Ya94,Pa93,Ha96,Fo98}, of oscillation searches at short baseline \cite{Ma96} and long baseline \cite{Ya94,Ha96,Ii93} accelerator facilities, and of double beta decay \cite{Kl99}. Given the phenomenological equivalence of violations of Lorentz invariance (VLI) \cite{Co97} and of the equivalence principle (VEP) \cite{Gl97}, neutrino oscillation searches can be generally interpreted as tests of fundamental principles of both special and general relativity, with a sensitivity at levels below $10^{-20}$ (see, e.g., \cite{Co98,Ha96}). Here we focus on the case of VEP or VLI induced oscillations coexisting with standard oscillations. Mixed scenarios of this kind have been considered, e.g., in \cite{Ga89,Co98,Ha96,Mi95} but, to our knowledge, they have not been discussed on the basis of Super-Kamiokande atmospheric $\nu$ observations so far. In the presence of several concurrent processes leading to $\nu_\mu\leftrightarrow\nu_\tau$ oscillations, the global Hamiltonian $H$ is the sum of several $2\times 2$ matrices $H_n$, which can be diagonalized through separate rotations (with angles $\xi_n$) of the flavor basis $(\nu_\mu,\nu_\tau)$. As far as models of the kind $\lambda^{-1}\propto E^{n}$ are concerned, the rotation angle $\xi$, which diagonalizes the total hamiltonian, is related to the oscillation length $\lambda$ through equations of the form \cite{Co98} \begin{mathletters} \begin{eqnarray} \pi\lambda^{-1}\sin2\xi &=& \left| \sum_n c_n\, \sin 2\xi_n\, E^n\, e^{i\eta_n}\right|\ , \label{lambdacsia}\\ \pi\lambda^{-1}\cos2\xi &=& \sum_n c_n\, \cos 2\xi_n\, E^n\ , \label{lambdacsib} \end{eqnarray} \end{mathletters} where the coefficients $c_n$ parametrize the strength of each oscillation mechanism. In general, only one of the complex phase factors $e^{i\eta_n}$ can be rotated away, the others being physically observable \cite{Co98,Ha96}. For any given choice of the parameters $(c_n,\xi_n,\eta_n)$, one has to derive the values of $\lambda$ and $\xi$ from the previous equations, and insert them in Eq.~(\ref{P}) to get the flavor transition probability. In the specific case of standard+VEP $(n=-1\oplus n=+1)$ oscillations, Eqs.~(\ref{lambdacsia}) and (\ref{lambdacsib}) can be rewritten as \cite{Ha96} \begin{mathletters} \begin{eqnarray} \pi\,\lambda^{-1}\,\sin{2\xi} &=& \left| 1.27 \, \frac{\Delta m^2}{E}\, \sin2\psi + 5.07\, \frac{|\phi|\Delta\gamma}{10^{-21}}\,E \, \sin{2\theta_G}\,e^{i\eta} \right|\ , \label{pilambdaa}\\ \pi\,\lambda^{-1}\,\cos{2\xi} &=& 1.27 \, \frac{\Delta m^2}{E}\, \cos2\psi + 5.07\, \frac{|\phi|\Delta\gamma}{10^{-21}}\,E \, \cos{2\theta_G} \ , \label{pilambdab} \end{eqnarray} \label{pilambda} \end{mathletters} where the following units have been used: $[\Delta m^2]=10^{-3}$ eV$^2$, $[L]=[\lambda]=10^3$ km, and $[E]= {\rm GeV}$. The same equations formally apply to violations of Lorentz invariance, modulo the replacements $|\phi|\Delta\gamma\to\delta v/2$ and $\theta_G\to\theta_v$ \cite{Gl97}. Notice that the oscillation phase $\pi\lambda^{-1}L$, to be inserted in Eq.~(\ref{P}), is proportional to the geometric average of the right hand sides of the above equations, so it will contain, besides the standard term $(\propto E^{-1})$ and the VEP term $(\propto E)$, also an energy-independent interference term. The oscillation amplitude $\sin^2 2\xi$ also acquires a nontrivial dependence on the neutrino energy through Eq.~(\ref{pilambda}). \subsection{Constraints from Super-Kamiokande data} Before performing a detailed fit to the SK data in the standard+VEP oscillation scenario, let us derive some qualitative bounds on the magnitude of possible VEP terms. According to the conclusions of Sec.~II, we expect that the second (VEP) term on the right hand of Eqs.~(\ref{pilambdaa}) and (\ref{pilambdab}) should be typically much smaller than the first (standard) term, namely \begin{equation} 5.07\,\frac{|\phi|\Delta\gamma}{10^{-21}}\,\frac{E}{\rm GeV} \ll 1.27\, \frac{\Delta m^2}{10^{-3}{\rm\ eV}^2}\,\frac{\rm GeV}{E}\ , \end{equation} or equivalently (for $\Delta m^2\sim {\rm\ few}\times 10^{-3}$ eV$^2$): \begin{equation} |\phi|\Delta\gamma \ll 10^{-21} \left( \frac{\rm GeV}{E}\right)^2\ . \label{e2} \end{equation} Therefore, the constraints to VEP effects should be stronger in the highest-energy SK data sample (UP$\mu$ events). Since the parent neutrino energy spectrum for UP$\mu$'s is peaked around $10^2$ GeV, the sensitivity to VEP-induced oscillations is expected to reach levels of $O(10^{-25})$ in the parameter $|\phi|\Delta\gamma$ (or, equivalently, in the parameter $\delta v/2$ for the VLI case). Of course, such sensitivity depends somewhat on $\theta_G$. In order to get some insight about the $\theta_G$-dependence of the expected constraints, let us consider the extreme values $\theta_G=0$ and $\theta_G=\pi/4$, and take $(\Delta m^2,\sin^2 2\psi)$ at their best-fit values $(2.8\times 10^{-3}{\rm\ eV}^2,1)$. We also fix $e^{i\eta}=1$ for simplicity. Then $P_{\mu\tau}$ takes a simple form, \begin{equation} P_{\mu\tau}=\left\{ \begin{array}{ll} \displaystyle\frac{1}{1+x^2}\sin^2\left(A\sqrt{1+x^2}\right) &,\ \theta_G=0\ , \\ \\ \sin^2\Big(A(1+x)\Big) &,\ \theta_G=\pi/4\ , \end{array} \right. \label{px} \end{equation} where $A=3.56\,L/E$, $x=1.42\,E^2|\phi|\Delta\gamma/10^{-21}$, and the units are $[E]={\rm GeV}$ and $[L]=10^3{\rm\ km}$. The standard case ($P^{\rm std}_{\mu\tau}=\sin^2A$) is recovered for $x=0$. In the UP$\mu$ event sample, where the VEP effect is larger, the value of $A$ is typically small $(\lesssim 0.6)$, and one can easily check numerically [from Eq.~(\ref{px})] that the difference $P_{\mu\tau}-P^{\rm std}_{\mu\tau}$ grows more rapidly with $x$ for $\theta_G=\pi/4$ than for $\theta_G=0$. Therefore, we expect a higher sensitivity to $x$ (i.e., to $|\phi|\Delta\gamma$) at larger $\theta_G$. Moreover, at small values of both $A$ and $x$ it turns out that $P_{\mu\tau}-P^{\rm std}_{\mu\tau}<0$ ($>0$) for $\theta_G=0$ ($\theta_G=\pi/4$), so that the muon rates should be less (more) suppressed than in the standard oscillation case. Figure~5 illustrates quantitatively the previous considerations, by showing the VEP effect (added to best-fit standard oscillations) on the SK muon distributions for two representative cases: (i) $\theta_G=0$ and $|\phi|\Delta\gamma=1.5\times 10^{-24}$ (solid line, blue in color); (ii) $\theta_G=\pi/4$ and $|\phi|\Delta\gamma=2.0\times 10^{-26}$ (dotted line, green in color). We anticipate that the such values are close to the border of the parameter region excluded by SK. The standard oscillation curves ($|\phi|\Delta\gamma=0$) are also shown for reference (dashed lines, red in color); they are identical to the best fit curves for standard oscillations ($n=-1$) in Fig.~2. As expected from the preceding discussion, the VEP effect is manifest at high energies (UP$\mu$ sample), and the expected deficit of UP$\mu$'s is more pronounced for $\theta_G=\pi/4$ than for $\theta_G=0$. The next step is to perform a $\chi^2$ analysis of the standard+VEP scenario. We take for the moment $(\Delta m^2,\sin^2 2\theta)=(2.8\times 10^{-3}{\rm\ eV}^2,\,1)$ and $e^{i\eta}=1$, while leaving the parameters ($|\phi|\Delta\gamma, \theta_G$) free. We find an important result, that strengthens the conclusions of Sec.~II: The $\chi^2$ fit is never improved in the presence of VEP effects, as compared with the value $\chi^2_{\min}=20.3$ derived in Fig.~2 for the standard case. Thus, not only VEP-induced oscillations cannot be the {\em leading\/} mechanism underlying the SK observations, but also there is no indication in favor of {\em subleading\/} VEP oscillation terms. As a consequence, we can place well-defined upper bounds on violations of relativity principles in the $(\nu_\mu,\nu_\tau)$ sector. Figure~6 shows the 90\% and 99\% C.L. limits on $|\phi|\Delta\gamma$ as a function of the VEP mixing parameter $\sin^22\theta_G$. The same limits apply to the neutrino asymptotic speed difference $\delta v/2$, as a function of the VLI mixing parameter $\sin^22\theta_v$. As expected from the discussion at the beginning of this subsection, the limits obtained in Fig.~6 are roughly of $O(10^{-25})$, and become stronger as $\theta_G$ increases. The stringent bounds in Fig.~6, together with the results shown in Fig.~3, represent our main contribution to the current understanding of atmospheric neutrino oscillations induced by violations of relativity principles in the $(\nu_\mu,\nu_\tau)$ sector. Notice that such bounds pre-empt the region of VEP (or VLI) parameters explorable with proposed long-baseline accelerator neutrino facilities \cite{Ha96}.% \footnote{Notice that the values of $L$, $E$, and $L\cdot E$ probed in the UP$\mu$ sample by Super-Kamiokande are higher than in proposed long-baseline neutrino beams.} Finally, we have investigated the robustness of the bounds shown in Fig.~6 under variations of the standard mass-mixing parameters. We have repeated the fit by varying $\Delta m^2$ and $\sin^2 2\psi$ within the 90\% C.L. limits shown in Fig.~4, and also by taking negative values for $\Delta\gamma$, as well as generic values for the complex phase $\eta$. We have found that, in any case, the value of $\chi^2_{\min}$ is not smaller than in the best-fit standard case. Therefore, standard+VEP (or standard+VLI) oscillations never represent a better description of the SK data, as compared to best-fit standard oscillations. Under the above variations, the upper bounds shown in Fig.~6 are somewhat modified within factors of a few, but do not change qualitatively: they always become stronger as $\sin^2 2\theta_G$ increases. The most conservative upper bound (including negative $\Delta \gamma$ cases) turns out to be \begin{equation} |\phi\Delta\gamma| < 3 \times 10^{-24} {\rm\ \ at\ 90\%\ C.L.}\ , \end{equation} independently of $\theta_G$. In the specific case $\theta_G=\pi/4$, the above bound can be lowered at least to $\lesssim 10^{-25}$. Analogous limits apply to the VLI parameters $|\delta v|/2$ and $\theta_v$. In particular, \begin{equation} |\delta v| < 6 \times 10^{-24} {\rm\ \ at\ 90\%\ C.L.}\ \end{equation} To our knowledge, the above limits to violations of special or general relativity principles are the strongest ever placed in the $(\nu_\mu,\nu_\tau)$ sector. They are valid under an assumption which is supported by the present data and appears likely to be corroborated in the future, namely, that standard $\nu_\mu\leftrightarrow\nu_\tau$ oscillations generated by $\nu$ mass and mixing represent the dominant mechanism underlying the Super-Kamiokande observations. \section{Summary and Conclusions} Among the $\nu_\mu\leftrightarrow\nu_\tau$ models with oscillation length following a power-law energy dependence, standard neutrino oscillations generated by $\nu$ mass and mixing are unique in providing a good description of the Super-Kamiokande atmospheric neutrino data, and are strongly favored as leading mechanism for $(\nu_\mu,\nu_\tau)$ flavor transitions. Additional, subleading $\nu_\mu\leftrightarrow\nu_\tau$ oscillations generated by possible violations of special or general relativity in the neutrino sector do not improve the agreement with the data, and must thus have a relatively small (or zero) amplitude. In particular, the fractional difference of asymptotic $\nu$ velocities $|\delta v|/2$ or of $\nu$ couplings to gravity $|\phi\Delta\gamma|$ cannot exceed the value $\sim 3\times 10^{-24}$ at 90\% C.L.\ for unconstrained neutrino mixing. The broadness of the neutrino energy range probed by Super-Kamiokande is crucial to obtain such strong limits. \acknowledgements We thank M.\ Gasperini for inspiring discussions and for useful comments. The work of A.M.\ and G.S.\ is supported by the Italian Ministero dell'Universit\`a e della Ricerca Scientifica e Tecnologica through a PhD grant.
\section{Introduction} There has been considerable study, over the last decade, of the frustrated spin-$\case 1/2$ square lattice Heisenberg antiferromagnet (the ``$J_1\mbox{-}J_2$ antiferromagnet"). These studies include exact diagonalizations on small systems\cite{dag89,sin90,sch92,ric93}, spin-wave calculations\cite{cec92,dot94}, series expansions\cite{gel89,Isingexp}, and a field-theoretic large-$N$ expansions\cite{read91}. These studies, and others, have provided a substantial body of evidence that the ground state of this system, in the region $0.4 \lesssim J_2/J_1 \lesssim 0.6$, has no long-range magnetic order and has a gap to spin excitations. For $J_2/J_1 \lesssim 0.4$ the model has conventional antiferromagnetic N\'eel order whereas for $J_2/J_1 \gtrsim 0.6$ the system orders in a columnar $(\pi,0)$ phase. Whether this ``intermediate phase" is a spatially homogeneous spin-liquid, or whether it has some type of spontaneously broken symmetry leading to a more subtle type of long-range order, has not been conclusively established. Zhitomirsky and Ueda\cite{zhi96} have proposed a plaquette resonating valence bond (RVB) phase, which breaks translational symmetry along both $x$ and $y$ axes, but preserves the symmetry of interchange of the two axes. The horizontal and vertical dimers resonate within a plaquette. An early series study\cite{gel90} had investigated the relative stability of various spontaneously dimerized states and had concluded that a columnar dimerized phase was the most promising candidate for the intermediate region, in agreement with the large-$N$ expansions. Zhitomirsky and Ueda\cite{zhi96} claim their plaquette phase has a lower energy than this columnar dimer phase, but we find this to be incorrect. Further support for the columnar dimer scenario comes from recent work of Kotov {\it et al.}\cite{dimerexp}, who combine an analytic many body theory with extended series and diagonalization results to study the nature and stability of the excitations in the intermediate region. It is argued that where the N\'eel phase becomes unstable the system will develop not only a gap for triplet excitations but also a gapped low-energy singlet which reflects the spontaneous symmetry breaking. This is clearly seen in the calculations. At $J_2/J_1 \simeq 0.38$ a second order transition occurs, with the energies of N\'eel phase and dimerized phase joining smoothly, and the energy gap and dimerization vanishing. It is the aim of this paper to further investigate, using series methods, the competing possibilities of columnar dimerization versus plaquette order in the intermediate region of the $J_1\mbox{-}J_2$ antiferromagnet. It is conceivable that both occur, with a transition from one to the other. However, such a transition, reflecting a change of symmetry, is expected to be first-order and not well suited to series methods. If both phases are locally stable the most direct way to compare them is by comparison of the ground state energies. If one is unstable this should show up by the closing of an appropriate gap or by the divergence of an appropriate susceptibility. In this paper we calculate the ground state energy and the singlet and triplet excitation spectra by series expansions about a disconnected plaquette Hamiltonian. We also calculate the susceptibility for the dimer phase to break translational symmetry in the direction perpendicular to the dimers. This susceptibility will be large if there is substantial resonance in the dimer phase and will diverge if there is an instability to the plaquette RVB phase. Combining the plaquette expansion results with the dimer expansions of Kotov {\it et al.}\cite{dimerexp}, a very interesting picture emerges for the quantum disordered phase. We find that the plaquette phase is unstable and hence is not the ground state for this model. The dimer phase, on the other hand, is stable. However, there is substantial resonance in the dimer phase. The spin-spin correlations are not simply those of isolated dimers. Instead, the nearest neighbor correlations are nearly identical along the rungs and chains of dimer columns. In contrast, the correlations from one dimer column to the next are much weaker. The spin-gap phase appears separated from the N\'eel phase by a second order transition, whereas it is separated from the columnar phase by a first order transition. These results are in remarkable agreement with the large-N theories \cite{read91}. The existence of a quantum critical point separating an antiferromagnetic phase and a quantum disordered phase with striped correlations in a microscopic model makes this critical point a particularly interesting one. The role of doping and its implications for high-$T_c$ materials deserves further attention. \section{Series Expansions and Results} We study the Hamiltonian \begin{equation} H = J_1 \sum_{{\rm n.n.}} {\bf S}_i\cdot {\bf S}_j + J_2 \sum_{{\rm n.n.n.}} {\bf S}_i\cdot {\bf S}_j \label{H} \end{equation} where the first sum runs over the nearest neighbor and the second over the second nearest neighbor spin pairs of the square-lattice. We denote the ratio of couplings as $y= J_2/J_1$. The linked-cluster expansion method has been previously reviewed in several articles\cite{he90,gel,gelmk}, and will not be repeated here. To carry out the series expansion about the disconnected-plaquette state for this system, we take the interactions denoted by the thick solid and dashed bonds in Fig. 1 as the unperturbed Hamiltonian, and the rest of the interactions as a perturbation. That is, we define the following Hamiltonian \begin{equation} H = H_0 + H_1 \end{equation} where the unperturbed Hamiltonian ($H_0$) and perturbation ($H_1$) are \begin{eqnarray} H_0 &=& J_1 \sum_{\langle ij\rangle \in A} {\bf S}_i \cdot {\bf S}_j + J_2 \sum_{\langle ij\rangle \in B} {\bf S}_i\cdot {\bf S}_j \nonumber \\ & & \\ H_1 &=& \lambda J_1 \sum_{\langle ij\rangle \in C} {\bf S}_i\cdot {\bf S}_j + \lambda J_2 \sum_{\langle ij\rangle \in D} {\bf S}_i\cdot {\bf S}_j \nonumber \end{eqnarray} and the summations are over intra-plaquette nearest-neighbor bonds (A), intra-plaquette second nearest-neighbor bonds (B), inter-plaquette nearest-neighbor bonds (C), inter-plaquette second nearest-neighbor bonds (D), shown in Fig. 1. With this Hamiltonian, one can carry out an expansion in powers of $\lambda$, and at $\lambda=1$ one recovers the original Hamiltonian in Eq. (\ref{H}). Thus, although we expand about a particular state, i.e. a plaquette state, our results at $\lambda=1$ describe the original system without broken symmetries, provided no intervening singularity is present. Such perturbation expansions about an unperturbed plaquette Hamiltonian have been used previously to study Heisenberg models for CaV$_4$O$_9$\cite{cavoser}. It is instructive to consider the states of an isolated plaquette. There are two singlet states, one with energy $(-2+y/2) J_1$ and the other with energy $(-3y/2) J_1$. The former is the ground state for $y<1$ and corresponds to pair singlets resonating between the vertical and horizontal bonds of the plaquette. It is even under a $\pi/2$ rotation. The latter is the ground state for $y>1$ and is odd under a $\pi/2$ rotation. The wavefunctions for these two singlet states are \begin{eqnarray} \psi_1 &=& {1\over \sqrt{12}} {\Big [} {\Big (} \begin{array}{cc} + & + \\ - & - \end{array} {\Big )} + {\Big (} \begin{array}{cc} + & - \\ + & - \end{array} {\Big )} + {\Big (} \begin{array}{cc} - & - \\ + & + \end{array} {\Big )} + {\Big (} \begin{array}{cc} - & + \\ - & + \end{array} {\Big )} -2 {\Big (} \begin{array}{cc} + & - \\ - & + \end{array} {\Big )} -2 {\Big (} \begin{array}{cc} - & + \\ + & - \end{array} {\Big )} {\Big ]} \nonumber \\ &=& {1\over \sqrt{3}} {\Big [} {\Big (} \begin{array}{c} \begin{picture}(30,10)(-5,-5) \put(0,0) {\circle{6}} \put(3,0){\line(1,0){12}} \put(18,0){\circle{6}} \end{picture} \\ \begin{picture}(30,10)(-5,-5) \put(0,0) {\circle{6}} \put(3,0){\line(1,0){12}} \put(18,0){\circle{6}} \end{picture} \end{array} {\Big )} + {\Big (} \begin{array}{cc} \begin{picture}(10,30)(-5,-5) \put(0,0) {\circle{6}} \put(0,3){\line(0,1){12}} \put(0,18){\circle{6}} \end{picture} & \begin{picture}(10,30)(-5,-5) \put(0,0) {\circle{6}} \put(0,3){\line(0,1){12}} \put(0,18){\circle{6}} \end{picture} \end{array} {\Big )} {\Big ]} \\ \psi_2 &=& {1\over 2} {\Big [} {\Big (} \begin{array}{cc} + & + \\ - & - \end{array} {\Big )} - {\Big (} \begin{array}{cc} + & - \\ + & - \end{array} {\Big )} + {\Big (} \begin{array}{cc} - & - \\ + & + \end{array} {\Big )} - {\Big (} \begin{array}{cc} - & + \\ - & + \end{array} {\Big )} {\Big ]} \nonumber \\ &=& {\Big [} {\Big (} \begin{array}{cc} \begin{picture}(10,30)(-5,-5) \put(0,0) {\circle{6}} \put(0,3){\line(0,1){12}} \put(0,18){\circle{6}} \end{picture} & \begin{picture}(10,30)(-5,-5) \put(0,0) {\circle{6}} \put(0,3){\line(0,1){12}} \put(0,18){\circle{6}} \end{picture} \end{array} {\Big )} - {\Big (} \begin{array}{c} \begin{picture}(30,10)(-5,-5) \put(0,0) {\circle{6}} \put(3,0){\line(1,0){12}} \put(18,0){\circle{6}} \end{picture} \\ \begin{picture}(30,10)(-5,-5) \put(0,0) {\circle{6}} \put(3,0){\line(1,0){12}} \put(18,0){\circle{6}} \end{picture} \end{array} {\Big )} {\Big ]} \end{eqnarray} where \begin{picture}(25,6)(-3,-3) \put(0,0) {\circle{6}} \put(3,0){\line(1,0){12}} \put(18,0){\circle{6}} \end{picture} means these two spins form a singlet. There are three triplet states, one with energy $(-1+y/2) J_1$ and a degenerate pair with energy $(-y/2) J_1$; like the singlets, these have a level crossing at $y=1$. Under a $\pi/2$ rotation the former is odd, while the latter two are even and odd, respectively. Finally there is a quintuplet state at $(1+y/2) J_1$, which is even under a $\pi/2$ rotation. For $y<1/2$ and $y>2$ the first excited state of the plaquette is a triplet, while for $1/2<y<2$ it is the other singlet. These states and corresponding energies are shown in Figure 2. The eigenstates of $H_0$, the unperturbed Hamiltonian, are direct products of these plaquette states. To derive the plaquette expansions we identify each plaquette as a 16 state quantum object, and these lie at the sites of a square lattice with spacing $2a$, where $a$ is the original lattice spacing. Interactions between plaquettes connect first and second-neighbor sites on this new lattice. The cluster data is thus identical to that used by us previously\cite{Isingexp} to derive Ising expansions for this model. Because there are 16 states at each cluster site, the vector space grows very rapidly with the number of sites and thus limits the maximum attainable order for plaquette expansions to considerably less than can be achieved for dimer or Ising expansions. We have computed the ground state energy $E_0$ to order $\lambda^7$, for fixed values of the coupling ratio $y$. The series are analysed using integrated differential approximants\cite{gut}, evaluated at $\lambda=1$ to give the ground state energy of the original Hamiltonian. The estimates with error bars representing confidence limits, are shown in Figure 3. For comparison we also show previous results obtained from Ising expansions\cite{Isingexp} and dimer expansions\cite{dimerexp}. We find that, in the intermediate region, the ground state energy for both plaquette and dimer phase are very close to each other and cannot be used to distinguish between them. The dimer expansion yields slightly lower energies near the transition to the N\'eel phase. We do not draw any conclusions from this. Zhitomirsky and Ueda\cite{zhi96} have claimed that the ground state energy from a second-order plaquette expansion is -0.63 (at $y=0.5$), much lower that the dimer expansion result -0.492. This result appears incorrect. At $J_2/J_1=\case 1/2$ the ground state energy is given by \begin{eqnarray} 4 E_0/NJ_1 &=& -7/4 - 277 \lambda^2/1440 -0.001357 \lambda^3 -0.0210609 \lambda^4 \nonumber \\ && -0.000319586 \lambda^5 -0.00580643 \lambda^6 -0.001822686 \lambda^7 +O(\lambda^8) \end{eqnarray} The second order result (at $\lambda=1$) is $E_0/NJ_1 = -0.485$, rather than $-0.63$. We note that if the second order coefficient were 4 times larger then the resulting energy would be $-0.62986$. We have also derived series, to order $\lambda^6$, for the singlet and triplet excitation energies, $\Delta_s(k_x,k_y)$, $\Delta_t(k_x,k_y)$ using the method of Gelfand\cite{gel}, and taking as unperturbed eigenfunctions the corresponding plaquette states. The low order terms for $J_2/J_1=0.5$ are given by: \begin{eqnarray} \Delta_s(k_x,k_y)/J_1 &=& 1 -301\lambda^2/1440 + 137 \lambda^3/86400 + 217 \lambda^3 \cos(k_x) \cos(k_y) /172800\nonumber \\ && +(-5\lambda^2/16 - 89 \lambda^3/9600 ) [\cos(k_x)+\cos(k_y)]/2 \\ \Delta_t(k_x,k_y)/J_1 &=& 1 - 3691 \lambda^2/30240 + (- 2 \lambda/3 + 11 \lambda^2/720 ) [\cos(k_x)+\cos(k_y)]/2 \nonumber \\ && -\lambda^2 [\cos(2 k_x)+\cos(2 k_y)]/120 + (\lambda/3 -5\lambda^2/96 ) \cos(k_x) \cos(k_y) \nonumber \\ && - \lambda^2 [\cos( 2 k_x) \cos( k_y) + \cos(k_x) \cos( 2k_y)]/90 + 7 \lambda^2 \cos( 2 k_x) \cos( 2 k_y)/360 \end{eqnarray} The full series are available on request. We first consider the triplet excitations. Figure 4 shows $\Delta_t (k_x,k_y)$ along high symmetry directions in the Brillouin zone for $\lambda=0.5$ and various coupling ratios $y$. For $\lambda \lesssim 0.6$ the series are well converged and direct summation and integrated differential approximants give essentially identical results. We find that the minimum gap occurs at $(0,0)$ for $J_2/J_1 \lesssim 0.55$ and moves to $(\pi,0)$ for $J_2/J_1\gtrsim 0.55$. Next we seek to locate the critical point $\lambda_c$ where the triplet gap vanishes. This is done using Dlog Pad\'e approximants to the gap series at the appropriate $(k_x,k_y)$. In practice this works well when the minimum gap lies at $(0,0)$. For $J_2=0$ we find a critical point at $\lambda_c=0.555(10)$. We can compare this result with recent work of Koga {\it et al.}\cite{koga98} who obtain $\lambda_c = 0.112$ from a modified spin-wave theory and $\lambda_c\simeq 0.54$ from a 4th order plaquette expansion. The critical point $\lambda_c$ increases with increasing $y$. At $y=0.5$, at the approximate centre of the intermediate phase, we find $\lambda_c \simeq 0.89(7)$. This result has some uncertainty but, if accurate, means that the plaquette phase becomes unstable before the full Hamiltonian ($\lambda=1$) is reached. The associated critical exponent $\nu$ describing the vanishing of the triplet gap is about 0.7 for $J_2/J_1 < 0.4$, suggesting that the transition lies in the universality class of the classical $d=3$ Heisenberg model. On the other hand, for $J_2/J_1 \gtrsim 0.4$ the exponent $\nu$ is about 0.4. This supports the existence of an intermediate phase lying in a different universality class. Figure 5 shows the singlet excitation energy $\Delta_s (k_x,k_y)$ along high symmetry directions in the Brillouin zone for $\lambda=0.5$ and the same coupling ratios $y$ as Figure 4. Again the series are well converged and direct summation and integrated differential approximants give essentially identical results. We find that the minimum gap occurs at $(0,0)$ for all $J_2/J_1$. We have also noted that for $J_2/J_1=0.5$, the triplet excitation and the singlet excitation have same gap at $\lambda=0$, but at $\lambda=0.5$, the singlet gap is considerable larger than the triplet gap, this means probably that the triplet gap close before the singlet gap at $J_2/J_1=0.5$. The critical point obtained by the Dlog Pad\'e approximant to the singlet gap is also generally slightly larger than that obtained from the triplet gap around $J_2/J_1=0.5$ (see Fig. 6). The full phase diagram in the parameter space of $J_2/J_1$ and $\lambda$ could be very interesting from the point of view of quantum phase transitions, but may not be easy to determine by numerical methods. Some possible scenarios are shown in Fig. 7. One possibility is that the plaquette phase, for all $J_2/J_1$, has an instability to some magnetic phase and the dimerized phase exists only very close to $\lambda=1$ inside the magnetic phases. A second possibility is that the plaquette-N\'eel critical line meets the N\'eel-dimer critical line at some multicritical point at a value of $J_2/J_1$ around $0.5$, after which there is a first order transition between the plaquette and the dimer phases. A third possibility is that the plaquette-N\'eel, N\'eel-dimer and plaquette-columnar critical lines all meet at some multicritical point. The numerically determined phase diagram is particularly uncertain in the interesting region, $0.5\lesssim J_2/J_1\lesssim 0.6$., where incommensurate correlations could also become important. Lastly we have derived expansions for a number of generalized susceptibilities. These are defined by adding an appropriate field term \begin{equation} \Delta H = h \sum_{ij} Q_{ij} \end{equation} to the Hamiltonian and computing the susceptibility from \begin{equation} \chi_Q = - {1\over N} \lim_{h\to 0} {\partial^2 E_0(h) \over \partial h^2} \end{equation} A divergence of any susceptibility signals an instability of that phase with respect to the particular type of order incorporated in $\chi$. We have computed two different susceptibilities from the plaquette expansion. One is the antiferromagnetic (N\'eel) susceptibility with the operator $Q_{ij}$ \begin{equation} Q_{i,j} = (-1)^{i+j} S^z_{i,j} \end{equation} The other is the dimerization susceptibility with the operator $Q_{ij}$ \begin{equation} Q_{i,j} = {\bf S}_{i,j}\cdot {\bf S}_{i+1,j} - {\bf S}_{i,j}\cdot {\bf S}_{i,j+1} \end{equation} which breaks the symmetry of interchange of $x$ and $y$ axes. We have computed series to order $\lambda^5$ for the antiferromagnetic susceptibility and to order $\lambda^4$ for the dimerization susceptibility. The series have been analyzed by Dlog Pad\'e approximants. The series for the antiferromagnetic susceptibility shows the same critical points (within error bars) as those obtained from the triplet gap for $J_2/J_1 \lesssim 0.4$. The series for the dimerization susceptibility is very irregular, and does not yield useful results. For example, for $J_2/J_1=0.5$, the series is: \begin{equation} \chi_d = 629/90 + 101\lambda /300 + 2.0097647 \lambda^2 -0.269629 \lambda^3 + 0.438527 \lambda^4 +O(\lambda^5) \end{equation} For completeness, we also compute the susceptibility for the dimer phase to become unstable to the plaquette phase from an expansion about isolated columnar dimers\cite{dimerexp}, by adding the following field term: \begin{equation} \Delta H = h \sum_{i,j} (-1)^j {\bf S}_{i,j}\cdot {\bf S}_{i,j+1} \end{equation} which breaks the translational symmetry in the direction perpendicular to the dimers. The series has been computed up to order $\lambda^7$, (note that $\lambda$ here is the parameter of dimerization). An analysis of the series shows that this susceptibility becomes very large as $\lambda\to 1$, for all $J_2/J_1$ and the critical $\lambda$, where the susceptibility appears to diverge, approaches unity from above as $J_2/J_1$ is increased to $0.5$. This implies that there are staggered bond correlations in the direction perpendicular to the dimers, which extend over a substantial range. An interesting question is, in the absence of the plaquette phase as discussed earlier, what could these correlations represent? At this stage it is useful to recall another calculation by Kotov {\it et al.}\cite{dimerexp}. Within the dimer expansion, they calculated two different dimer order parameters, \begin{equation} D_x = |<{\bf S}_{i,j}\cdot {\bf S}_{i+1,j}>- <{\bf S}_{i+1,j}\cdot {\bf S}_{i+2,j}>|, \end{equation} and, \begin{equation} D_y = |<{\bf S}_{i,j}\cdot {\bf S}_{i+1,j}>- <{\bf S}_{i,j}\cdot {\bf S}_{i,j+1}>|, \end{equation} where the elementary dimers connect spins at ${i,j}$ and ${i+1,j}$. They found that for $0.4\lesssim J_2/J_1\lesssim0.5$, $D_y$ is nearly zero, whereas $D_x$ only goes to zero at the critical point. These results suggest that the dimer phase consists of strongly correlated two-chain ladders, which are then weakly correlated from one ladder to next. This striped nature of spin correlations in the dimer phase has not been noted before and is clearly a very interesting result. The situation for $J_2/J_1\gtrsim 0.5$ is again less clear. As discussed before, there are many possibilities for the phase diagram in that region and much longer series are needed to throw more light on the situation. Perhaps there is an interesting multicritical point in that region of the phase diagram. \section{Discussion} We have attempted to further elucidate the nature of the intermediate, magnetically disordered, phase of the spin-$\case 1/2$ $J_1\mbox{-}J_2$ Heisenberg antiferromagnet on the square lattice. This phase is believed to occur in the range $0.4 \lesssim J_2/J_1 \lesssim 0.6$. Our approach has been to derive perturbation expansions (up to order $\lambda^7$) for the ground state energy, singlet and triplet excitation energies, and various susceptibilities, starting from a system of decoupled plaquettes ($\lambda=0$) and extrapolating to the homogeneous lattice ($\lambda=1$). We have also derived expansions about an unperturbed state of isolated columnar dimers (``dimer phase"). Both of these have been proposed as candidates for the intermediate phase. We find that the ground state energy for both plaquette and dimer phases are very similar, any difference lying within the error bars. From this result alone we cannot favor one phase over the other. The analysis of the singlet and triplet excitation spectra suggests an instability in the plaquette phase. In particular, in the disconnected plaquette expansions, Dlog Pad\'e analysis indicates that the gaps would vanish for $\lambda$ less than unity. The gap appears to close first for the triplets and then for the singlets. This is the strongest evidence that the plaquette phase is {\it not} realized in this model. However, we should mention here that the critical exponents associated with the vanishing of the gaps are rather small ($<0.4$) and the gap closes not too far from $\lambda$ equal to unity. Thus, with a relatively short series, this should be treated with some caution. One could ask why the energy series appear to converge well despite the instability. However, this is a well known feature of series expansions, that quantities having weak singularities may continue to show reasonable values even if extrapolated past the singularity. A consistent interpretation of these results is that within the parameter space of our non-uniform Hamiltonian, the plaquette phase is first unstable to a magnetic phase, which then must give way to the columnar dimer phase. Similar results for the instability of the staggered dimer phase were suggested before by Gelfand et al\cite{gel89}. However, the full phase diagram in the $J_2/J_1$ and $\lambda$ parameter space is difficult to obtain reliably, especially near the transition to the columnar phase. There are possibilities of some novel multicritical points, which deserve further attention. One of our most interesting results is the finding of striped spin correlations in the dimer phase. In this phase, the nearest neighbor spin correlations are nearly equal along the rungs and along the chains of a two spin column and there are extended bond correlations along the chains. However, spin correlations from one column to the next are much weaker. In other words, the dimers are strongly resonating along vertical columns. The existence of a quantum critical point separating an antiferromagnetic phase with such a quantum disordered phase with striped correlations is a very interesting feature of this model which deserves further attention in the context of high-$T_c$ materials. \acknowledgments We would like to thank Subir Sachdev and Oleg Sushkov for many useful discussions. This work has been supported in part by a grant from the National Science Foundation (DMR-9616574) (R.R.P.S.), the Gordon Godfrey Bequest for Theoretical Physics at the University of New South Wales, and by the Australian Research Council (Z.W., C.J.H. and J.O.). The computation has been performed on Silicon Graphics Power Challenge and Convex machines. We thank the New South Wales Centre for Parallel Computing for facilities and assistance with the calculations.
\section{Introduction} The Major Atmospheric Gamma-ray Imaging Cherenkov (MAGIC) Telescope will be a large imaging air Cherenkov telescope for ground-based $\gamma$-ray observations above 10\,GeV. Details of its design, the scientific motivation, the feasibility of the project, and other issues have been described elsewhere (Barrio et al. \cite{barrio}). The telescope is expected to become operational by the middle of 2001. The telescope is optimised to achieve the lowest energy threshold and highest flux sensitivity achievable with present technology. This makes it applicable to a large range of astrophysical research fields: \begin{itemize} \item Blazars (study of EGRET blazars, possible discovery of additional sources) \item Cosmology (measurements of the near-infrared background via $\gamma$-$\gamma$ absorption) \item Investigation of the pulsed $\gamma$-ray emission from pulsars \item Search for gamma-emission from Supernova remnants (and hence search for evidence of production and acceleration of cosmic rays) \item Search for decay/annihilation line-emission from WIMPs clustering at the galactic centre \item Identification of ``unidentified EGRET sources'' (position accuracy $\approx$ 1.2') \item Search for high-energy counterparts of gamma-ray bursts \end{itemize} The last part of this scientific program will briefly be discussed in this article. \section{Fast follow-up observations} In order to make fast follow-up observations of GRBs possible, the MAGIC Telescope will have an interface (probably a socket connection) to the GRB Coordinate Network GCN (Barthelmy et al. \cite{barthelmy} and these proceedings) and a secondary ground station for direct communication with HETE II (Ricker et al., these proceedings). HETE II is a dedicated GRB research satellite which is expected to be launched in 2000. It will provide approx. 30 burst positions per year with accuracies better than 10 arcmin. These notifications are expected to arrive with a delay of less than 5\,s. The MAGIC Telescope is specially designed to have low inertia such that the telescope can be positioned on any point in the sky within less than 30\,s. In case of a notification, a fast check of the observability of the GRB location will be performed. If the decision is positive, the present observations will be stopped immediately and a special fast drive will position the telescope on the GRB which will then be observed for the remaining night-time and probably also the following night in order to detect possible delayed emission. Taking into account a decision time of 5 s at our site, we expect the reaction time between the actual start of the burst and the start of the follow-up observation to be $30\pm10$ s. The MAGIC Telescope will thus be able to perform observations of non-delayed emission of all bursts with durations above 30\,s. This is the position of the rightmost peak in the BATSE burst duration ($T_{90}$) distribution (see e.g. Kouveliotou et al. \cite{kou}) and corresponds to a fraction $\approx$ 33\,\% of all bursts which trigger BATSE on the 64 ms time scale. The telescope will be built on the Canary Islands (Tenerife or La Palma). Cherenkov telescopes can only observe during the night. Observations will be possible up to zenith angles of $\approx 80^\circ$, i.e. about 40 \% of the total sky ($4\pi$) will be accessible. Assuming 30 \% of the nights to have bad weather and taking into account that the presence of the moon can prevent observations of certain positions, we arrive at a duty cycle of $\approx$ 10 \%. The effective field of view of the photo-sensor camera of the MAGIC Telescope is 1.6$^\circ$ in diameter. We will therefore be able to safely observe any of the positions provided by HETE II and also some from the GCN and expect $\approx$ 5 serious immediate follow-up observations per year. For delayed emission (time-window of the order of an hour up to several days) this number will be larger. \section{Expected performance} The expected performance in terms of sensitivity is summarised in Figure \ref{fig-magflusens}. This is the performance we expect for the second phase of the project in which high quantum efficiency hybrid photo sensors will replace the photo-multipliers of the telescope's camera. In the first phase our threshold will be $\approx$ 30 GeV, in the second phase 10 GeV. It is not yet clear how soon the second phase will follow the first. The sensitivity at energies $>$ 30\,GeV is to a good approximation independent of these changes. \begin{figure} \leavevmode \centering \epsfxsize=8.8cm \epsffile{grbfluence.eps} \caption{\label{fig-magflusens} The fluence sensitivity of the MAGIC Telescope (MAGICT) compared to other experiments (MAGICT refers to observations at zenith angles up to $\approx$ 30$^\circ$, ``MAGICT large ZA'' to those at $\approx$ 70$^\circ$). The fluence sensitivity at threshold $E_0$ is defined here as the total fluence above $E_0$ which is necessary for a $5 \sigma$ detection of the burst above that threshold. A total burst duration of 60 s and a differential spectral index of 2.6 is assumed. For MAGICT the actual observation time is assumed to be only 30 s taking into account the reaction time. For all other detectors the observation time is assumed to be equal to the burst duration.} \end{figure} Given the diverse shape of GRB light-curves it is difficult to predict an average counting rate for gamma-rays in successful burst observations. We note instead that the MAGIC Telescope will have an effective collection area for primary gamma photons of $10^8$\,cm$^2$ at the threshold rising to $10^9$\,cm$^2$ at 100 GeV. For a hypothetical counting rate of 0.1\,Hz for EGRET above 100 MeV, we expect (assuming a spectral index of 2.0) a counting rate of $\approx$ 100\,Hz above 10 GeV. A rate of $\approx$ 6 Hz for 30 s will suffice for a 5 $\sigma$ detection with a moderate background rejection applied. For strong bursts such as GRB930131, gamma counting rates of the order of 1 kHz may occur. The data acquisition will therefore be prepared to sustain such rates without additional dead-time. \section{Conclusions} The MAGIC Telescope will be able to make a major contribution to GRB research by providing high sensitivity measurements in the (for GRBs) essentially unexplored regime above 10 GeV. The high counting rates in strong bursts will also permit to study the shape of the light-curve in more detail than previously possible (time resolutions of the order of 1 s). Search for small delays with respect to the low energy emission is thus possible. As an interesting side-result this may provide one of the best possible lower limits to the quantum gravity energy scale as pointed out by Amelino-Camelia et al. (\cite{amelino}).
\section{Introduction} When non magnetic sites are diluted in an unfrustrated ferromagnet with a probability $\mu$, the transition temperature is reduced and vanishes at the percolation threshold $1-\mu_{P,0}$. In such non frustrated systems, two phases only exist: a low temperature ferromagnetic phase above the percolation threshold, and a paramagnetic phase. If the temperature is decreased at the percolation threshold, the dynamics becomes slower because large-scale droplet-like objects of size $\xi_T$ form, with $\ln{\xi_T} \sim J/T$~\cite{Coniglio}. These objects have energy barriers scaling like the logarithm of their volume~\cite{Rammal,Henley}. This results in a slow dynamics and interrupted aging~\cite{RM-tree} ({\sl i.e.} with a finite relaxation time~\cite{Bouchaud}). This shows that despite the absence of frustration, the simplest models of dilute magnets already have a phenomenology close to the one of spin glasses, even though freezing in these systems is a cross-over due to an increasing correlation length becoming of order of the system size. This indicates that some perturbations of these unfrustrated systems may drive them to a true spin glass phase, which we show in the present article by studying the thermodynamics of a particular model. In dilute magnet compounds, such as Eu$_x$Sr$_{1-x}$S~\cite{EuSrS,note,exp}, a low temperature spin glass phase appears close to the percolation threshold. The main features of the phase diagram are: \begin{itemize} \item[(i)] As the dilution $\mu$ is increased from the pure system with $\mu=0$, the ordering temperature decreases. \item[(ii)] A tricritical point exists at a dilution $\mu_t$ and temperature $T_t$, with $1-\mu_t$ of order of the percolation threshold $1-\mu_{P,0}$ in the absence of frustration. At this tricritical point, the ferromagnetic, paramagnetic and spin glass phases meet. \item[(iii)] As dilution is increased from $\mu_t$, the spin glass transition temperature decreases from $T_t$ at the tricritical point to zero at the percolation threshold $1-\mu_P$, with $1-\mu_P$ the percolation threshold of the system with frustration, smaller than the percolation threshold $1-\mu_{P,0}$ in the absence of frustration. \item[(iv)] The spin glass phase is reentrant inside the ferromagnetic phase. \end{itemize} One purpose of the present article is to show that these qualitative features of the phase diagram can be reproduced in a model that combines dilution and short range frustration. This model does not consist in a detailed microscopic modeling of Eu$_x$Sr$_{1-x}$S, but rather contains the generic ingredients entering the physics of these systems (dilution and short range frustration)~\cite{Rammal-Souletie}. Even though this treatment relies on a specific lattice topology (a tree structure), Bethe-Peierls phase diagrams are equivalent to mean field phase diagrams while the Bethe-Peierls method is powerful enough to give an exact answer to the issue of reentrance. The resulting phase diagrams are therefore not expected to be specific to our treatment but are generic features of the coupling Hamiltonian. The article is organized as follows. The model is given in section~\ref{sec:model}. The paramagnetic phase boundary is solved in section~\ref{sec:recur-rel}. We study in section~\ref{sec:reentrance} the spin glass~/~ferromagnet phase boundary and show that it is reentrant. The issue of reentrance is non trivial because Nishimori's argument~\cite{Nishimori,Kitatani,Georges-LeDoussal,LeDoussal-Harris,Horiguchi} does not hold in our model, part of the exchanges being frozen. An extension of the model is given in which the spin glass~/~ferromagnet phase boundary is shown not to reenter in the ferromagnetic phase asymptotically close to the tricritical point, whereas it has a turning point at a lower temperature. We expect this unusual type of phase diagram to be a generic feature of models combining disordered and frozen exchanges, and may be obtained in finite dimensional models also. Finally, we study in section \ref{sec:dil-spinliquid} the effect of diluting a spin liquid state, in which case ordering by disorder generates a transition from an Ising paramagnet to an Ising spin glass. This ordering by disorder mechanism is ``minimal'' because spins have an Ising symmetry only. \section{The model} \label{sec:model} \subsection{The dilute magnet model and its generalization} In a Bethe-Peierls calculation, only the properties of the ``top'' spin (the highest one in the hierarchy; see Fig.~\ref{fig:husimi}) are considered, and the thermodynamic limit is obtained by growing the number of generations to infinity~\cite{Bethe,Peierls,Baxter}. The top spin fixed point magnetization distributions $P^*(m)$ are of three types: (i) {\it paramagnetic phase}: $P^*(m) = \delta(m)$; (ii) {\it spin glass phase}: $P^*(m)$ is even; (iii) {\it ferromagnetic phase}: $P^*(m)$ has a finite first moment. The transitions between the phases are of a mean field type~\cite{Bethe,Peierls,Baxter}. The phase diagram of the $\pm J$ model on the Bethe lattice~\cite{spinglass1,spinglass2,spinglass3,spinglass4, spinglass5,Carlson,spinglass6} is very similar to the one of the Sherrington-Kirkpatrick model~\cite{SK}, and the Bethe-Peierls treatment allows a correct description~\cite{spinglass6} of the Almeida-Thouless line~\cite{Almeida-Thouless}. This shows that a Bethe-Peierls treatment succeeds in reproducing the mean field phase diagram of spin glass models. We consider a model in which the ferromagnetic bonds $J$ of the Cayley tree are canceled with a probability $\mu$. Frustration is added by completing the triangles, with frozen antiferromagnetic bonds $\tau$ forming a Husimi cactus-like structure~\cite{Thorpe} (see Fig.~\ref{fig:husimi}). We use the binary variables $\theta_i=0,1$, with $J_i = \theta_i J$. The temperature is expressed in units of $J$. The Hamiltonian is \begin{equation} H = - \sum_{\langle i,j \rangle} \theta_{i,j} \sigma_i \sigma_j + \tau \sum_{\langle i,j \rangle'} \sigma_i \sigma_j , \end{equation} where $\langle i,j \rangle$ denotes the bonds of the tree structure and $\langle i,j \rangle'$ the next nearest neighbor pairs of sites in the same generation. The distribution of the $\theta$-variable is \begin{equation} \label{eq:bondinitial} p(\theta) = (1-\mu) \delta(\theta-1) + \mu \delta(\theta) . \end{equation} \begin{figure} \centerline{\psfig{file=Figure1.eps,width=8cm}} \caption{The Husimi cactus-like structure model of dilute magnet with frustration. The structure with a top spin $x$ is obtained by gluing the two structures with top spins $y$ and $z$. The top spin is the highest one in the hierarchy (at site $x$ on this figure). The ferromagnetic bonds of the tree are canceled with a probability $\mu$ and a fixed antiferromagnetic coupling $\tau$ is added. } \label{fig:husimi} \end{figure} The specificity that some bonds are frozen in this model has drastic consequences on the shape of the phase diagram, as we will show. The Husimi cactus structure allows the introduction of a {\sl local} frustration resulting from next-nearest-neighbor interactions, and can be used to mimic the effects of local antiferromagnetic interactions in dilute compounds. Chandra and Dou\c{c}ot~\cite{Doucot} considered a frustrated spin model on a regular Husimi cactus structure, and studied ordering by disorder in the spin liquid state in the Bethe-Peierls limit (see also~\cite{Quantum} for a study of the effect of quantum fluctuations). These authors considered a non disordered model in which a spin glass phase cannot exist~\cite{Doucot}. In our model with randomness, we show the stability of a spin glass solution in some regions of the phase diagram. We consider bond instead of site percolation because the site percolation threshold of the Husimi cactus structure would be equal to the one of the tree structure, independent of the additional bonds $\tau$. The bond percolation model is therefore better suited for modeling dilute compounds~\cite{EuSrS}, since the bond percolation threshold of the structure without frustration (a tree structure) is $1-\mu_{P,0}=1/2$, larger than the bond percolation threshold $1-\mu_P =1-1/\sqrt{2}$ of the structure with frustration (the Husimi cactus structure shown on Fig.~\ref{fig:husimi}). For the sake of generality, we not only consider a dilute magnet model with $\theta=0,1$, but extend the bond distribution~(\ref{eq:bondinitial}) to incorporate possible antiferromagnetic bonds on the tree structure. The distribution of the bond variables $\theta$ is \begin{equation} \label{eq:bond} p(\theta) = (1-\lambda)(1-\mu) \delta( \theta - 1) + \mu \delta(\theta) + \lambda (1-\mu) \delta(\theta + 1) , \end{equation} while the additional antiferromagnetic bonds $\tau$ are frozen. This model interpolates between the dilute model with a short-range frustration $\tau$ ($\lambda=0$), and the $\pm J$ model ($\mu=0$ and $\tau=0$). \subsection{Absence of a Nishimori line argument} Our Hamiltonian is formally invariant under local gauge transformations $\sigma_i \rightarrow \epsilon_i \sigma_i$, $J_{i,j} \rightarrow \epsilon_i \epsilon_j \sigma_i \sigma_j$, with $\epsilon_i=\pm 1$~\cite{Toulouse}. In some spin glass models (such as the $\pm J$ model), gauge invariance provides strong constraints on the phase diagram. The internal energy can be calculated exactly on Nishimori's line by expanding the average energy over the gauge group~\cite{Nishimori}. This line crosses the phase boundary at the tricritical point~\cite{Nishimori,Kitatani,Georges-LeDoussal,LeDoussal-Harris}. Moreover, spin correlations can be related to gauge variable correlations, with the consequence that the frontier between the ferromagnetic and spin glass phases is either vertical or reentrant~\cite{Nishimori}. In our model, gauge invariance is useless for the following reason. One can define a local distribution of bond variables $P_{i,j}(J)$, being (\ref{eq:bond}) on the tree bonds, and $\delta(J+\tau)$ on the antiferromagnetic bonds $\tau$. Nishimori's line is defined by $ \beta_N = \frac{1}{2} \ln{\left(\frac{1-\lambda}{\lambda} \right)} $~\cite{Horiguchi}, and $\beta_N = + \infty$. The first equality originates from the tree bond variables and the second from the frozen antiferromagnetic bonds $\tau$. The two equalities can be formally met if $\lambda=0$, in which case Nishimori line is $\beta_N = + \infty$. However, this does not make predictions on the phase diagram possible even in the case $\lambda=0$~\cite{Note} because one does not expect to be able to describe finite temperature spin glass properties in terms of the ground state only (the only state selected if $\beta_N=+\infty$). Nishimori's argument can therefore not be made in this model, and the question of reentrance (item (iv) in the introductory section) cannot be answered on the basis of Nishimori's line while Bethe-Peierls calculations are powerful enough to allow the derivation of exact results. We show that the spin glass~/~paramagnet boundary is reentrant in the dilution model ($\lambda=0$) in the $(\mu,T)$ plane. In the $\pm J$ model with the additional coupling $\tau$ ($\mu=0$), and in the $(\lambda,T)$ plane, we show that the spin glass~/~ferromagnet phase boundary is not reentrant asymptotically close to the tricritical point, whereas it has a turning point at lower temperatures. This behavior, richer than in usual spin glass models, is to our opinion a generic feature of Hamiltonians combining disorder and frozen bonds. We conjecture the existence of finite dimensional models with a similar phase diagram. \section{Recursion relations} \label{sec:recur-rel} We now derive the recursion of the top-site magnetization when cacti are glued as shown on Fig.~\ref{fig:husimi}. We denote by $m_x$ the magnetization at site $x$, and $m_y$ and $m_z$ the magnetizations at the descendant sites $y$ and $z$. The derivation of the recursion relations in our dilute magnet model is similar to the case of the $\pm J$ model (see Ref.~\cite{Carlson}). \subsection{Recursions and the paramagnet phase boundary} \label{sec:recursion} Following Ref.~\cite{Carlson}, we denote by ${\cal Z}^{(\pm)}_x$ the conditional partition function with the spin at site $x$ frozen in the direction $\pm$. The magnetization at site $x$ is $ m_x = ({\cal Z}^{(+)}_x - {\cal Z}^{(-)}_x) /({\cal Z}^{(+)}_x + {\cal Z}^{(-)}_x) . $ The partition functions ${\cal Z}^{(\pm)}_x$ are related to the partition functions ${\cal Z}^{(\pm)}_{y,z}$ of the descendant sites according to \begin{equation} \label{eq:Z} {\cal Z}^{(\sigma_x)}_x = \sum_{\sigma_y,\sigma_z} W^B_{\theta_y,\theta_z}(\sigma_x|\sigma_y,\sigma_z) {\cal Z}_y^{(\sigma_y)} {\cal Z}_z^{(\sigma_z)} , \end{equation} with the Boltzmann weight factor \begin{equation} \label{eq:WB} W^B_{\theta_y,\theta_z}(\sigma_x|\sigma_y,\sigma_z) = \exp{\left( \beta ( \theta_y \sigma_x \sigma_y + \theta_z \sigma_x \sigma_z ) \right)} \exp{\left( -\beta \tau \sigma_y \sigma_z \right)} , \end{equation} and $\theta=0, \pm 1$. We next trace over the spins at sites $y$ and $z$ in Eq.~\ref{eq:Z} to obtain \begin{equation} \label{eq:recursion} m_x = f(m_y,m_z|\theta_y,\theta_z) = p \frac{ m_y (\theta_y - u \theta_z) + m_z (\theta_z - u \theta_y)} {1 - u p^2 \theta_y \theta_z + m_y m_z (p^2 \theta_y \theta_z -u)} , \end{equation} with $p= \tanh{(\beta J)}$ and $u = \tanh{(\beta \tau)}$. The recursion of the magnetization distribution is \begin{equation} \label{eq:PM} P_{n+1}(m_x) = \int d m_x d m_y \sum_{\theta_y,\theta_z} p(\theta_y) p(\theta_z) P_n(m_y) P_n(m_z) \delta\left(m_x - f(m_y,m_z|\theta_y,\theta_z) \right) , \end{equation} with $p(\theta)$ the distribution of bond variables (\ref{eq:bond}), and $P_n$ the magnetization distribution of the top spin with $n$ levels of hierarchy. We denote by $\langle \langle m^k \rangle \rangle_n$ the moment of order $k$ of $P_n(m)$. We now parametrize the tricritical line, where the three phases (paramagnetic, ferromagnetic and spin glass) meet. The meeting point of these phases is a {\sl line} in the parameter space $(\lambda, \mu ,T)$. If $\lambda$ [$\mu$] is fixed and the phase diagram is considered in the $(\mu,T)$ [$(\lambda,T)$] plane, the three phases meet in a tricritical {\sl point}. Let us first consider the stability of the paramagnetic solution with respect to perturbations in the first moment. To lowest order the recursion of the first moment is $ \label{eq:recur-m} \langle \langle m \rangle \rangle_{n+1} = 2 p \langle \langle G_{y,z} \rangle \rangle \langle \langle m \rangle \rangle_n $, with $ G_{y,z} = (\theta_y - u \theta_z)/(1 - u p^2 \theta_y \theta_z) $. The disorder average of $G$ is understood as $ \langle \langle G_{y,z} \rangle \rangle = \sum_{\theta_y,\theta_z} p(\theta_y) p(\theta_z) G_{y,z} $. The paramagnetic solution is stable with respect to perturbations in the first moment if $2 p \langle \langle G_{y,z} \rangle \rangle < 1$. A similar reasoning shows that the paramagnetic solution is stable with respect to perturbations in the second moment if $2 p^2 \langle \langle G_{y,z}^2 \rangle \rangle <1$. To summarize, the tricritical line is defined by \begin{equation} \label{eq:tricritical} 2 p \langle \langle G_{y,z} \rangle \rangle = 1 \mbox{ , and } 2 p^2 \langle \langle G_{y,z}^2 \rangle \rangle = 1 . \end{equation} \subsection{Limiting cases} \label{sec:limiting} \begin{figure} \centerline{\psfig{file=Figure3.ps,height=8cm}} \caption{Phase diagram of the pure Husimi cactus system (all the coupling $\theta$ being unity) as a function of the frustration $\tau$. A low-temperature ferromagnetic phase is present if $\tau<1$. If $\tau>1$, the system is a spin liquid (it does not order even at zero temperature).} \label{fig:tau} \end{figure} \noindent {\sl The pure system:} Let us consider the recursion (\ref{eq:recursion}) in the pure system limit in which the variables $\theta$ are all equal to unity. This amounts to specializing the distribution (\ref{eq:bond}) to the case $\lambda=\mu=1$ while keeping finite the local frustration $\tau$. The only possible phases are ferromagnetic and paramagnetic. The recursion of the magnetization is $ m_{n+1} = 2 p (1-u) m_n/ (1 - p u^2 + (p^2 - u)m_n^2) . $ The paramagnetic phase is stable against ferromagnetic fluctuations if $ 2 p (1-u)/(1 - p u^2) < 1 $. The phase diagram is shown on Fig.~\ref{fig:tau} as a function of the frustration $\tau$ with a spin liquid phase if $\tau > 1$. When considering in the following the frustrated magnet model, we assume $\tau < 1$, in which case the pure system has an ordered phase at low temperature. Diluting the spin liquid state with $\tau>1$ is examined in section~\ref{sec:dil-spinliquid}. \begin{figure} \centerline{\psfig{file=Figure4.ps,height=8cm}} \caption{Phase diagram of the dilute magnet model with frustration ($\lambda=0$, $\tau=0.1$). The spin glass phase exists below the percolating dilution $\mu_P =2^{-1/2} \simeq 0.707$. The paramagnetic~/~spin glass phase boundary inside the ferromagnetic phase is unphysical, as well as the paramagnetic~/~spin glass boundary inside the spin glass phase. The solid line is obtained from the calculation in section~\ref{sec:dilutemodel} of the frontier between the spin glass and ferromagnetic phases. This solution is exact close to the critical point and we have continued it to lower temperatures by an arbitrary linear behavior. The exact zero temperature spin glass~/~ferromagnet phase boundary is $\mu_0 \simeq 0.24191$. } \label{fig:para} \end{figure} \bigskip \noindent {\sl $\pm J$ model:} The $\pm J$ model is recovered if $\mu=\tau=0$, with a tricritical point at coordinates $p_t=1/2$, $\lambda_t=\frac{1}{2} \left(1 - \frac{1}{\sqrt{2}} \right)$~\cite{Carlson}. \bigskip \noindent {\sl Dilute magnet with frustration:} If $\lambda=0$ and $\tau \ne 0$ the relations (\ref{eq:tricritical}) become \begin{eqnarray} \label{eq:p-f} &&\frac{2 p (1 - \mu)(1-u)(1 - \mu u p^2)}{1 - u p^2} = 1\\ &&\frac{2(1-\mu)p^2}{(1-u p^2)^2} \left( (1-\mu)(1-u)^2 + \mu (1+u^2) (1 - u p^2)^2 \right) = 1, \label{eq:p-sg} \end{eqnarray} determining the paramagnetic phase boundary shown on Fig.~\ref{fig:para}. The frontier between the spin glass and ferromagnetic phases will be examined in Section~\ref{sec:reentrance} by looking for an instability in the first moment of the spin glass solution. The zero temperature limit of the paramagnetic/spin glass phase boundary can be obtained by considering first the limit $p=1$ in Eq. (\ref{eq:p-sg}), and second the limit $u=1$. The order in which the two limits are taken is imposed by the fact that $1-p \ll 1-u \ll 1$ at low temperatures because $\tau<1$. One finds the limit to be $1-\mu = 1-1/\sqrt{2}$. As it is expected, this dilution is equal to the percolation threshold of the Husimi cactus structure. \section{Frontier between the spin glass and ferromagnetic phases} \label{sec:reentrance} \subsection{Method} In order to determine the frontier between the spin glass and ferromagnetic phases, we study the instability of the spin glass solution with respect to perturbations in the first moment. This involves first calculating the spin glass solution close to the tricritical line and next determining whether this solution is stable with respect to ferromagnetic fluctuations. Carlson {\it et al.}~\cite{Carlson} performed this calculation for the $\pm J$ model close to the tricritical point, and shown the spin glass phase to be marginally reentrant inside the ferromagnetic phase. By marginal, we mean that the spin glass~/~ferromagnet phase boundary has a quadratic behavior $\lambda_t - \lambda \sim (T_t - T)^2$, which is specific to this model. Other models (as the one we presently analyze) have a linear behavior $\lambda_t - \lambda \sim T_t - T$, with a positive (reentrant behavior) or negative prefactor (non reentrant behavior). The calculation follows Ref.~\cite{Carlson} where the $\pm J$ model was solved, and is asymptotically exact close to the tricritical point. This is complemented by an exact determination of the zero temperature phase boundaries which, to our knowledge, has not appeared previously in the literature even for the $\pm J$ model. We first determine the asymptotic spin glass solution close to the tricritical line. The second moment to lowest order is \begin{equation} \label{eq:2moment} \langle \langle m^2 \rangle \rangle = \frac{2p^2 \langle \langle G_{y,z}^2 \rangle \rangle -1}{4p_t^2 \langle \langle G_{y,z} G_{z,y} H \rangle \rangle_t} , \end{equation} with $ H = (p^2 \theta_y \theta_z - u)/ (1 - u p^2 \theta_y \theta_z) $, and the subscript ``t'' denoting a quantity evaluated on the tricritical line. We next consider a perturbation in the first moment of the spin glass solution. The recursion of the first moment is $ \langle \langle m \rangle \rangle_{n+1} = \kappa \langle \langle m \rangle \rangle_n$ , with $ \kappa=2p \langle \langle G_{y,z} \rangle \rangle -2p_t \langle \langle m^2 \rangle \rangle \langle \langle G_{y,z} H \rangle \rangle_t$ to order $(T_t - T)$. If $\kappa < 1$ the spin glass phase is stable and otherwise the ferromagnetic phase is stable. \subsection{$\pm J$ model with an additional short-range coupling $\tau$ -- Fixed $\mu$; ($\lambda$, T) phase diagram} \label{sec:pmJ} In the small-$\tau$ limit, it is straightforward to show that \begin{enumerate} \item{The tricritical point can be determined in an expansion in $u_t$: $ \label{eq:triperturb} p_t=1 - 2 \lambda_t =\frac{1}{\sqrt{2}}(1+\frac{u_t}{4}) $. } \item{The slope of the spin glass~/~ferromagnet phase boundary at the tricritical point is \begin{equation} \label{eq:dk-dT} \left. \frac{d \kappa}{d T} \right|_t = \frac{u_t}{4 T_t^0} \left( 1 - \frac{5 \sqrt{2}}{2 T_t^0} \right) \simeq -0.467 \, u_t \mbox{ } < 0 , \end{equation} with $T_t^0 = 1/ \tanh^{-1}{(1/\sqrt{2})} \simeq 1.135$ the tricritical point temperature with $\tau=0$. } \end{enumerate} >From what we deduce that the spin glass phase does not reenter inside the ferromagnetic phase close to the tricritical point. \begin{figure} \centerline{\psfig{file=Figure7.eps,width=8cm}} \caption{Possible shapes of the spin glass~/~ferromagnet phase boundary in the $\pm J$ model with a small antiferromagnetic coupling $\tau$ corresponding to a spin glass phase not reentrant inside the ferromagnetic phase asymptotically close to the tricritical point. This implies two possible behaviors: (a): no reentrance at any temperature (b): no reentrance close to the tricritical point but reentrance at lower temperatures. We prove that (b) is the correct behavior in a zero temperature exact solution. } \label{fig:shape} \end{figure} Notice that $\left. d \kappa / d T \right|_t$ in Eq. (\ref{eq:dk-dT}) vanishes if $\tau=0$. This is because reentrance is marginal in the $\pm J$ Bethe lattice spin glass~\cite{Carlson} and can therefore not be obtained from an expansion to first order in $T_t-T$. We have evaluated numerically the coefficient $\kappa$ with a finite $\tau$ and a finite $\mu$. As $\mu$ is increased above a critical value, the transition changes from reentrant to non reentrant. We now derive the exact spin glass solution in the zero temperature limit, which allows to discriminate rigorously between the two behaviors on Fig.~\ref{fig:shape} (a) and~\ref{fig:shape} (b). We look for the zero temperature fixed point spin glass and ferromagnetic solutions $P^*(m)$ under the form \begin{equation} \label{eq:P*} P^*(m) = \frac{x+y}{2} \delta(m-1) + (1-x) \delta(m) + \frac{x-y}{2} \delta(m+1) , \end{equation} with $x$ and $y$ the spin glass and ferromagnet order parameters. It turns out that the functional form of the magnetization distribution (\ref{eq:P*}) is stable when it is iterated in the zero temperature limit of (\ref{eq:recursion}) and (\ref{eq:PM}). To determine $x$ and $y$, we impose (\ref{eq:P*}) to be the fixed point magnetization distribution. The solution with a finite magnetization is $ x = (1 - 4 \lambda)(1 - 2 \lambda) $, and $y^2 = (1-4 \lambda)(1 - 8 \lambda)/ (1 - 2 \lambda)^2 $. Imposing $y^2>0$ leads to the intersection $\lambda_0=1/8=0.125$ of the spin glass~/~ferromagnet phase boundary and the zero temperature axis in the $(\lambda,T)$ plane. $\lambda_0$ is independent of the strength of the additional coupling $\tau$. If $\tau=0$, the value $\lambda_t$ of $\lambda$ at the tricritical point is $\lambda_t=\frac{1}{2}(1 - \frac{1}{\sqrt{2}}) \simeq 0.146$~\cite{Carlson}, larger than $\lambda_0$. The zero temperature solution is therefore consistent with the reentrant behavior of the spin glass~/~ferromagnet phase boundary of the $\pm J$ Bethe lattice spin glass~\cite{Carlson}. As $\tau$ is increased, the tricritical point $(\lambda_t(\tau), T_t(\tau))$ evolves with $\tau$ whereas the intersection of the spin glass~/~ferromagnet phase boundary and the zero temperature axis remains equal to $\lambda_0$. Therefore, if $\tau$ is small, $\lambda_t(\tau)$ remains larger than $\lambda_0$. From what we deduce the existence of a turning point in the spin glass~/~ferromagnet phase boundary (Fig.~\ref{fig:shape} (b)): the spin glass phase does not reenter close to the tricritical point whereas it reenters at lower temperatures. We believe this behavior to be generic of spin glass models with frozen exchanges and we conjecture that a similar behavior may be obtained in finite dimensional models. \subsection{Dilution with a short-range frustration $\tau$ -- $\lambda=0$; ($\mu$, T) phase diagram} \label{sec:dilutemodel} A small-$\tau$ perturbation calculation leads to $ p_t = 1 - \frac{1}{2} u_t $, $\mu_t = \frac{1}{2} - \frac{1}{2} u_t$, and $\left. d \kappa / d T \right|_t = \tau /2 T_t^2 \mbox{ } > 0$, which proves that the spin glass phase reenters in the ferromagnetic phase close to the tricritical point in the limit of small $\tau$. We have shown on Fig.~\ref{fig:para} the behavior of the spin glass~/~ferromagnet phase boundary. This phase boundary is exact only close to the tricritical point, and we have continued it by an arbitrary straight line at lower temperatures. The reentrant behavior is confirmed by zero temperature exact results. The paramagnetic~/~spin glass frontier intersects the zero temperature axis at the percolation threshold, and the spin glass~/~ferromagnet frontier intersects the zero temperature axis at $\mu_0$, the real root of $-10 \mu_0^3 + 6 \mu_0^2 - 5 \mu_0 + 1 = 0$, approximately $\mu_0 \simeq 0.24191$. This confirms the reentrant behavior of the spin glass transition in the dilute magnet model with frustration. \section{Diluting the spin liquid} \label{sec:dil-spinliquid} We now consider dilution in the regime $\tau>1$, {\sl i.e.} when the pure system is a spin liquid (see Fig.~\ref{fig:tau}). As it could be expected, a ferromagnetic instability of the spin liquid solution (Eq.~\ref{eq:p-f}) does not exist. However, a spin glass instability of the paramagnetic solution {\sl does} exist upon diluting the system. Let us first consider the zero temperature phases, in the limit $1-u \ll 1-p \ll 1$ (since $\tau>1$). The phase diagram at finite temperatures is shown on Fig.~\ref{fig:liquid}. A finite temperature spin glass phase opens from the point ($\mu=1/2$, $T=0$) as temperature is increased from zero. The low temperature phase boundary is $ T = 4(1-\tau) / [\ln{(\mu-1/2)^2}] $. This provides a simple situation in which diluting an Ising spin liquid results in an Ising spin glass phase. The underlying ordering by disorder mechanism~\cite{order} is analyzed in section~\ref{sec:ordering-disorder}. \section{Conclusion: diluting a frustrated magnet {\sl versus} diluting a spin liquid} We have shown the existence of a spin glass solution upon diluting both the weakly frustrated magnet ($\tau <1$), and the spin liquid ($\tau>1$). We underline the differences in the physics in these two regimes. \begin{figure} \centerline{\psfig{file=Figure9.ps,height=8cm}} \caption{Boundary between the spin glass and paramagnetic phases upon diluting the spin liquid state of our model ($\tau>1$). The spin glass phase is confined inside the boundary shown for $\tau=1.2$ ($\Diamond$), $\tau=1.25$ ($+$), and $\tau=1.3$ ($\Box$). The spin glass boundary collapses onto the point $\mu=1/2$ in the zero temperature limit. This boundary behaves like $T \sim 1 / \ln{| \mu - 1/2|}$ around this singular point The spin glass phase is favored upon increasing the temperature (ordering by disorder -- see section~\ref{sec:ordering-disorder}). } \label{fig:liquid} \end{figure} \subsection{Diluting the frustrated ferromagnet} We believe the generation of a spin glass phase upon weakly frustrating a dilute magnet close to the percolation threshold to be due to the following: {\sl the strong diluted unfrustrated magnet is already close to a spin glass}. This can be seen on the example of square lattice dilute magnets~\cite{Rammal}, where dilution removes sites in the ferromagnet up to the point where the percolating cluster is a fractal object at the percolation point. Since the order of ramification of percolating clusters is finite~\cite{ramification}, one can isolate large droplet-like objects~\cite{droplets} from the remaining of the structure by cutting a finite number of bonds. This results in large sets of spins that can be reversed at a finite energy cost, thus being responsible for the existence of quasi-degenerate ground states separated by a large distance in phase space (with different magnetizations~\cite{RM-fractal}), and with barriers scaling like the logarithm of their volume~\cite{Rammal}. The addition of frustration in dilute magnets close to the percolation threshold turns the quasi spin glass order into a true one. We have shown this explicitly in our model, and a similar behavior was obtained in another model~\cite{Georges-LeDoussal}. We do not expect the low energy states of the spin glass phase with a small frustration to be very different from the droplet-like states of the unfrustrated magnet. The ``chaos and memory'' behavior of metallic spin glasses was put forward in Ref.~\cite{Bouchaud-droplets}, associated to the growth of fractal droplets with a chaotic behavior in the sense that droplets at a given temperature overlap weakly with droplets at a different temperature. We do not expect a chaotic behavior in our model because the finite temperature droplet excitations of the frustrated dilute magnet should be obtained from reversing clusters of spins in the unfrustrated magnet dilute lattice. \subsection{Diluting the spin liquid: ordering by disorder} \label{sec:ordering-disorder} The mechanism for generating a spin glass phase from the spin liquid is different. The spin glass phase originates from a balance between the small-dilution regime in which {\sl dilution suppresses the liquid behavior in favor of spin-glass correlations}, and a large-dilution regime in which {\sl dilution suppresses spin glass correlations by cutting the system into finite pieces}. This is an order by disorder mechanism~\cite{order}: thermal fluctuations favor a spin glass arrangement and therefore reduce the phase space dimensionality compared to the one of the spin liquid state. Let us think in terms of low temperature properties in the large-$\tau$ limit. In this limit, the neighboring spins coupled by the strong antiferromagnetic exchange $\tau$ correlate antiferromagnetic, thus leaving mainly two residual degrees of freedom per bond $\tau$. We note $m_y$ and $m_z$ the magnetization of these two spins, and, for the sake of a qualitative argument, assume $m_y=-m_z$ as a result of the strong bond $\tau$. Let us assume the spins at sites $y$ and $z$ to be frozen and look whether freezing is relevant in the Bethe-Peierls limit. We see from Eq. (\ref{eq:recursion}) that $m_x=0$ if the two ferromagnetic bonds are present ({\sl i.e.} if the triangular plaquette is frustrated, $\theta_y=\theta_z=1$). In the unfrustrated plaquettes $\theta_y=0$, $\theta_z=1$, or $\theta_y=1$, $\theta_z=0$, correlations in the magnetization can propagate from one generation to the other. The system is cut into two pieces if $\theta_y=\theta_z=0$, preventing correlations to propagate from one generation to the other. When the ferromagnetic bonds are diluted, frustration is reduced since the fraction of frustrated triangular plaquettes $(1-\mu)^2$ decreases upon increasing the dilution $\mu$. Decreasing frustration therefore decreases the short range liquid-like correlations and favors a cooperative spin glass arrangement. In the large dilution limit, the exchanges are severely depleted and a paramagnetic behavior is restored since the system is cut into finite pieces. In between these two regimes, the unfrustred bond configurations $\theta_y=1, \theta_z=0$ and $\theta_y=0$, $\theta_z=1$ with a weight $2 \mu(1-\mu)$ dominate the physics and make a spin glass order possible. \subsection{Concluding remarks} Finally, we would like to compare the present work to other approaches developed previously in the literature, and mention some open questions. The phase diagram with {\sl all} the bonds drawn from the distribution (\ref{eq:bond}) was studied by Aharony~\cite{Aharony}, Giri and Stephen~\cite{Giri} and Viana and Bray~\cite{Viana-Bray}. These models share similarities with the dilute fcc antiferromagnets studied by de Seze~\cite{deSeze} and Wengel, Henley and Zippelius~\cite{Henley3}. Nieuwenhuizen and Nieuwenhuizen and van Duin studied the field theory of a model of site-disordered magnet~\cite{Nieuwen1,Nieuwen2}. One may also define a model similar to ours in a finite dimension. As we conjectured, a phase diagram similar to the one on Fig.~\ref{fig:shape} (b) may be obtained. On the other hand, it may be useful to investigate replica symmetry breaking in Bethe-Peierls calculations. Hierarchical lattices have been used previously by Georges and Le Doussal~\cite{Georges-LeDoussal} in relation with the renormalization group flow along Nishimori's line, and by Gingras and S{\o}rensen to study reentrance from a paramagnetic to a ferromagnetic phase~\cite{Gingras}. A model with frozen exchanges may be studied on a finite dimensional hierarchical lattice, which could be a first step in addressing the phase diagram on Fig.~\ref{fig:shape}(b) in a finite dimension. This approach should probably rely on a numerical iteration of the renormalization equations similar to Ref.~\cite{Gingras} while an analytic study was possible in the present work. Finally, ordering by disorder seems to be a generic behavior of spin liquids~\cite{Doucot,order,Quantum,Henley2,Simon}. We found in the present work an ordering by disorder resulting in a transition from a paramagnetic to a spin glass ordering in an Ising model. This may be viewed as a ``minimal'' ordering by disorder from a $Z_2$-symmetric paramagnet to a spin glass because the Ising order parameter has the lowest possible spin symmetry. \section*{Acknowledgments} The authors thank P. Chandra and B. Dou\c{c}ot who introduced one of us (R.M.) to Bethe-Peierls calculations. J. Souletie pointed out to us the existence of a spin glass phase upon diluting the spin liquid. P. Simon pointed out to us Ref.~\cite{Henley2}. P. Pujol pointed out to us Refs.~\cite{Nishimori,Kitatani,Georges-LeDoussal} and the absence of a Nishinori line argument. The authors also acknowledge fruitful discussions with A. Georges, P. Holdsworth, P. Le Doussal, M. Gingras, J.M. Maillard, H. Nishimori, and J. Villain. \newpage
\section{Introduction} The layered organic molecular crystals $\kappa $-(BEDT-TTF)$_{2}$X (where BEDT-TTF is bis-(ethylenedithia-tetrathiafulvalene) and X is an anion (e.g., X=I$_{3}$, Cu[N(CN)$_{2}$]Br, Cu(SCN)$_{2}$)) are particularly interesting because they are strongly correlated electron systems with similarities to the high-$T_{c}$ cuprate superconductors including unconventional metallic properties and competition between antiferromagnetism and superconductivity \cite{mck1,mck2,fuku,kanoda0}. Furthermore, they are available in high purity single crystals and, in contrast to the cuprates, their lower superconducting transition temperature ($T_{c}\sim 10$K) makes experimentally accessible in steady magnetic fields properties such as the upper critical field and Shubnikov-de Haas oscillations\cite{ish,wos}. Recently it has been argued that a minimal theoretical model that can describe these materials is a Hubbard model on an anisotropic triangular lattice with one hole per site\cite{mck2,fuku}. Calculations at the level of the random-phase approximation\cite{rpa} and the fluctuation-exchange approximation\cite{flex} suggest that at the boundary of the antiferromagnetic phase this model exhibits superconductivity mediated by spin fluctuations. As the anisotropy of the intersite hopping varies the model changes from the square lattice to the isotropic triangular lattice to decoupled chains\cite{mck2}. The wavevector associated with the antiferromagnetic spin fluctuations changes\cite{merino} and the superconductivity has been predicted to change from d-wave singlet (as in the cuprates) to s-wave triplet in the odd-frequency channel\cite{rpa}. Experimental results that are consistent with unconventional superconductivity include the temperature dependence of the nmr relaxation rate $1/T_{1}$ (including the absence of a Hebel-Slichter peak)\cite {desoto,kanoda}, the temperature and magnetic field dependence of the electronic specific heat\cite{nakazawa}, the temperature dependence of the thermal conductivity,\cite{belin} and the sensitivity of $T_{c}$ to disorder \cite{zuo}. The temperature dependence of the nmr Knight shift (which measures the electron spin susceptibility) in the superconducting state provides a means to distinguish triplet and singlet pairing. For triplet pairing the Knight shift does not change on entering the superconducting state, whereas for singlet pairing the Knight shift goes to zero as the temperature decreases to zero. The Knight shift of $^{13}$C nmr on the X=Cu[N(CN)$_2$]Br is consistent with the latter. In contrast, the Knight shift of $^{17}$O nmr on Sr$_2$RuO$_4$ is consistent with the former.\cite{ishida} If the superconductivity is spin singlet then the upper critical field cannot exceed the paramagnetic limit $H_{P}$, also known as the Pauli limit or Clogston-Shandrasekhar limit\cite{clogston,pauli}. Above $H_{P}$ the Cooper pairs are destroyed by the Zeeman splitting produced by the magnetic field coupling to the electronic spins. For weak-coupling BCS theory \begin{equation} H_{P} = H_{P}^{BCS} \simeq {\frac{1.8k_{B}T_{c} }{\mu _{B}}}. \label{bpauli} \end{equation} For $T_c = 10$ K, as in the material studied here, this gives $H_{P}^{BCS} = 18 $ T. Strong coupling effects\cite{perez} and d-wave pairing\cite{sondhi} only change this value of $H_{P}$ slightly. In most superconductors the paramagnetic limit is irrelevant because the superconductivity is destroyed at much lower fields due to the frustration of the orbital degrees of freedom associated with the formation of vortices. However, in layered superconductors with fields parallel to the layers the vortices can fit between the layers and paramagnetic limiting can become important.\cite {klemm} Previous determinations of the upper critical field of the $\kappa $% -(BEDT-TTF)$_{2}$X family\cite{oshima,kwok,lang,graebner,murata,delong} have mostly focussed on measurements of the slope ${\frac{dH_{c_{2}}(T)}{dT}} $ near $T_{c}$. The values obtained for X=Cu[N(CN)$_{2}$]Br and X=Cu(SCN)$_{2}$ are in the range 10 to 20 T/K. Using the WHH formula\cite{whh} for a three-dimensional superconductor, this very large slope would suggest a zero-temperature $H_{c_{2}}(T=0)=0.7T_{c}{\frac{dH_{c_{2}}(T)}{dT}=}$ 70-140 T, which is well above the BCS Pauli limit. A previous transport measurement on the X=Cu(SCN)$_{2}$ salt was carried out in pulsed magnet fields.\cite {saito} A quasi-linear temperature dependence was found with $H_{c2}\sim $25 T and the authors concluded that the upper critical field exceeded the Pauli limit. A study of the upper critical field of X=Cu(CN)[N(CN)$_2$]\cite {nakamura} determined from the resistive transition found an upper critical field of about 25 T for fields parallel to the layers. Studies on the lower T% $_{c}$ organic compounds such as the $\kappa $-(BEDT-TTF)$_{2}$I$_{3}$\cite {wanka}, $\beta $-(BEDT-TTF)$_{2}$I$_{3}$, and $\beta $-(BEDT-TTF)$_{2}$IBr$% _{2}$\cite{murata2} have found that the $H_{c2}$ at zero temperature lies below or close to the Pauli paramagnetic limit predicted by BCS theory. Similar paramagnetic field limited $H_{c2}$ have been reported in the cuprate YBa$_{2}$Cu$_{3}$O$_{7-\delta }$\cite{dzurak} and the heavy fermion superconductors UPd$_{2}$Al$_{3}$\cite{gloos}. If there is paramagnetic limiting there is theoretically the possibility that as the magnetic field is increased at low temperatures there is a first-order phase transition into non-uniform superconducting state, originally proposed by Fulde and Ferrell and Larkin and Ovchinikov\cite{ff}. As the dimensionality of the system decreases the magnetic field range over which this phase is stable increases.\cite{shimahara} Such a first-order phase transition was recently seen in ultrathin beryllium films.\cite{adams} It is still controversial about whether this phase does exist in UPd$_{2}$Al$% _{3}$\cite{gloos}. On the other hand, if the superconductivity is triplet there is also the possibility of re-entrant superconductivity at high fields such that $T_{c}(H)$ actually increases with increasing field\cite {dupuis,lebed2}. In this paper we report the measurement of the interlayer resistivity of X=Cu(SCN)$_{2}$ down to 0.5 K and up to 30 T for a range of field directions. For magnetic fields parallel to the layers, the upper critical field increases approximately linearly with decreasing temperature to values that clearly exceed the BCS Pauli limiting field (\ref{bpauli}), but are consistent with the paramagnetic limit, estimated directly from the superconducting condensation energy. The upper critical field as a function of angle shows a sharp cusp for fields almost parallel to the layers, consistent with two-dimensional decoupled layers. We find no evidence of a first order phase transition as a function of field at low temperatures. \section{Theoretical background} We now briefly summaries some theoretical results concerning the upper critical field which we will use later in interpreting our results. A more complete discussion can be found in Reference \onlinecite{bul}. \subsection{Angular dependence of the upper critical field} Anisotropic Ginzburg-Landau theory is valid when the coherence length perpendicular to the layers, $\xi _{\perp }$, is much larger than the interlayer spacing. It predicts that the dependence of the upper critical field on the angle $\theta $ between the field and the normal to the layers is\cite{klemm,bul,schneider} \begin{equation} \left[ \frac{H_{c2}(\theta )\cos (\theta )}{H_{c2\perp }}\right] ^{2}+\left[ \frac{H_{c2}(\theta )\sin (\theta )}{H_{c2\parallel }}\right] ^{2}=1, \label{anisot} \end{equation} where $H_{c2\perp }$ and $H_{c2\parallel }$ are the upper critical field for fields perpendicular and parallel to the layers, respectively. The perpendicular upper critical field is determined by $\xi _{\parallel }$, the coherence length parallel to the layers, \begin{equation} H_{c2\perp }={\frac{\Phi _{0}}{2\pi \xi _{\parallel }^{2}}} \label{hcperp} \end{equation} where $\Phi _{0}$ is the flux quantum. The coherence lengths parallel and perpendicular to the layers are related by \begin{equation} {\frac{\xi _{\parallel }}{\xi _{\perp }}}={\frac{H_{c2\perp }}{% H_{c2\parallel }}}. \label{ratio} \end{equation} Klemm, Luther, and Beasley considered the upper critical field of layered superconductors when the layers were infinitely thin.\cite{klemm} For both Lawrence-Doniach theory and microscopic theory, they found that for fields parallel to the layers, if the interlayer coupling is sufficiently weak the upper critical field diverges at low temperatures unless spin-orbit effects or paramagnetic limiting is present. This is because the Josephson vortices associated with the field parallel to the layers have no normal core and can fit between the layers. Bulaevskii\cite{bul} and Schneider and Schmidt\cite {schneider} considered a more general model where the layers have a finite thickness $d$, resulting in a finite upper critical field \begin{equation} {\frac{H_{c2\perp }}{H_{c2\parallel }}}={\frac{d}{\sqrt{12}\xi _{\parallel }}% }. \label{ratio2} \end{equation} They also found that if the coupling between the layers is sufficiently weak then the angular dependence of the upper critical field is given by \begin{equation} \left| \frac{H_{c2}(\theta )\cos (\theta )}{H_{c2\perp }}\right| +\left[ \frac{H_{c2}(\theta )\sin (\theta )}{H_{c2\parallel }}\right] ^{2}=1 \label{tink} \end{equation} This same angular dependence was found earlier for thin two-dimensional films by Tinkham using a simple fluxoid quantization argument.\cite{tinkham} The main difference from the anisotropic three-dimensional result is that at $\theta =$90$^{\text{o}}$, $H_{c2}(\theta )$ from Eqn. (\ref{anisot}) is smooth or bell-shaped with $\frac{dH_{c2}(\theta )}{d\theta }=0$ whereas $% H_{c2}(\theta )$ from Eqn. (\ref{tink}) has \ a cusp at $\theta =$90$^{\text{% o}}$. If the upper critical field is determined solely by coupling of the field to the spins then it will be independent of the field direction. Bulaevskii\cite {bul} considered the case where the paramagnetic limit is larger than the upper critical field for fields perpendicular to the layers but smaller than the upper critical field determined by orbital effects for fields parallel to the layers. The angular dependence is then given by \begin{equation} \left| \frac{H_{c2}(\theta )\cos (\theta )}{H_{c2\perp }}\right| \left[1- \left(\frac{H_{c2\perp}}{H_{c2\parallel }} \right)^2 \right] +\left[ \frac{% H_{c2}(\theta )}{H_{c2\parallel }}\right] ^{2}=1 \label{pauli-tink} \end{equation} where $H_{c2\parallel }=H_P$. This also results in an $H_{c2}$ versus $% \theta $ curve which has a cusp at $\theta =$90$^{\text{o}}$. Indeed the angular dependence is difficult to distinguish from Eqn. (\ref{tink}). \subsection{ Estimating the paramagnetic limiting field} The metallic phase has a finite Pauli spin susceptibility $\chi _{e}$ compared to the vanishing susceptibility (at zero temperature) of a spin singlet superconducting state. Hence, it will be energetically favorable to destroy the superconducting state when the magnetic energy density gained by the difference in susceptibilities exceeds the superconducting condensation energy density $U_{c}$. The critical field $H_{P}$ at which this occurs is given by\cite{clogston} \begin{equation} U_{c}={\frac{\mu _{0}}{2}}\chi _{e}H_{P}^{2} \label{cond} \end{equation} where $\mu _{0}$ is the magnetic permeability of free space. In BCS theory the condensation energy density is $U_c = {\frac{1 }{2}} N(E_F) \Delta(0)^2 $ where $N(E_F)$ is the metallic density of states and $% \Delta(0)=1.76 k_B T_c$ is the zero-temperature energy gap. Making using of these relations and $\chi_e = (\mu_B)^2 N(E_F)$ we obtain the expression (% \ref{bpauli}) for $H_P$. {\it Many-body effects}. In the $\kappa$-(BEDT-TTF)$_2$X crystals there are significant many-body effects; the electron effective mass $m^*$ determined from magnetic oscillations can be two to five times larger than that predicted by band structure calculations.\cite{mck2,wos} The effect of this on the paramagnetic limit needs to be taken into account. Perez-Gonzalez\cite {perez} finds that the paramagnetic limiting field is enhanced by a factor of $m^*/m_b$. However, he did not take into account the simultaneous effect on the Zeeman splitting: the $g$ factor changes to $g^*$. When this is done one finds that within a Fermi liquid framework the Pauli limit is actually reduced from (\ref{bpauli}) by a factor of $g^*/g$.\cite{mck0} This ratio can be estimated from thermodynamic measurements or from the spin-splitting of magnetic oscillations.\cite{mck0} The values obtained by these two methods for X=Cu(SCN)$_2$, are 0.8 and 1.4, respectively.\cite{mck0} Alternatively, we can make a {\it theory-independent} estimate of $H_P$ by using (\ref{cond}) and the experimentally determined condensation energy density and spin susceptibility. This method of determining $H_P$ is very attractive because it does include all the many-body effects (without assuming a Fermi liquid picture) and does {\it not} assume the validity of any particular theory of superconductivity for the material in question. Haddon et al.\cite{haddon} found $\chi_e = 4.3 \times 10^{-4}$ emu per mole (corresponding to a density of states of 7 states per (eV molecule)) for the X=Cu(SCN)$_2$ salt. By a reanalysis of Graebner et al.'s\cite{graebner} specific heat data Wosnitza\cite{wos} evaluated the condensation energy density in terms of the thermodynamic critical field $B_{th} = $ 90 mT where $U_c = {\frac{ 1 }{2 \mu_0}} B_{th}^2$. Taking the unit cell volume of 1695 $\AA^3$ and two (BEDT-TTF)$_2$X units in each unit cell gives $B_P = 30 \pm 5 $ T. The uncertainty is estimated based on the uncertainty in the values for the condensation energy and the susceptibility. \section{Experimental details} Single crystals of $\kappa $-(BEDT-TTF)$_{2}$Cu(SCN)$_{2}$ were synthesized by the electrocrystallization technique described elsewhere\cite{zuo}. The interlayer resistance was measured with use of the four probe technique. Contact of the gold wires to the sample was made with a Dupont conducting paste or graphite paste. Typical contact resistances between the gold wire and the sample were about 10 $\ \Omega $. A current of 1 $\mu A$ was used to ensure linear {\it I-V} characteristics. The voltage was detected with a lock-in amplifier at low frequencies of about 312 Hz. To avoid pressure effects due to solidification of grease, the sample was mechanically held by thin gold wires. The data presented in this work were taken in a $^{3}$He system with field up to 30 T at the National High Magnetic Field Laboratory at Tallahassee. The sample can be rotated in the field and the orientation was determined by using a Hall probe at low fields. \section{Results} Shown in Fig. 1 is a typical field dependence of the interlayer resistance plotted in a semi-log scale at a temperature of 4.2 K. The field is applied parallel to the planes. The resistive transition in parallel field is typical of the low dimensional organic superconductors with a broad transition width in field and a large positive magnetoresistance in the normal state. The superconducting transition or the upper critical field {\it H}$_{c2}$ is defined at the 1 $\Omega $ level. To check the validity of this criteria, the critical field will be compared with that obtained by a more conventional definition. Shown in the inset are the same data in a linear scale. The two lines are extrapolations of the normal state magnetoresistance and the superconducting transition with the upper critical field {\it H}$_{c2}${\it *} defined at the crossing point of the two lines. Fig. 2 is an overlay of resistive transitions in parallel field at different temperatures from T = 0.5 K to 10.2 K. With increasing temperature, the curves shift to the left toward lower critical fields. The transition curves are nearly parallel for all temperatures in the semilog scale. {\it H}$_{c2}$ is almost the mid-transition point as in a conventional superconductor, where parallel transitions are seen but in a linear scale. The temperature dependences of the two fields {\it H}$_{c2}$ and {\it H}$% _{c2}${\it *} are shown in Fig. 3. Within the scatter of the points, the two upper critical fields have nearly the same linear temperature dependence with {\it dH}$_{c2}${\it /dT} $\approx 3$ TK$^{-1}$. The offset in the superconducting transition temperature is due to the different definitions. The upper critical fields at zero temperature are about 30 T and 33 T for {\it H}$_{c2}$ and {\it H}$_{c2}${\it *,} respectively. The dashed line is the Pauli limit $H_{\text{{\it P}}}$ = 18.4 T, calculated from Eqn. (\ref {bpauli}) with{\it \ T}$_{c}$ = 10 K. Clearly, $H_{\text{{\it P}}}$ defined this way is well below the measured upper critical fields at low temperatures. On the other hand, $H_{c2}$ is consistent with our estimate of $H_{\text{{\it P}}}$ from thermodynamic quantities. To look at the anisotropy of the upper critical field, a systematic measurements have been taken as a function of angle $\theta ,$ defined between the field direction and the normal of the plane. Plotted in Fig. 4 is an overlay of resistive transitions as a function of field at different angles. The six curves are representative of the angular dependence from field parallel to the layers ($\theta =90^{o})$ to normal to the layers ($% \theta =180^{o})$. With increasing $\theta $, the field dependence of the resistive transition is drastically changed. At $\theta $ = 91.50$^{\text{o}% } $, {\it H}$_{c2}$ is decreased by about 4 T. At $\theta $ = 96.64$^{\text{o% }} $, a shoulder-like feature is developed in {\it R(H)} with a corresponding decrease in {\it H}$_{c2}$ by about 12 T. The shoulder-like structure develops into a well defined peak at $\theta $ = 178$^{\text{o}}$ with the occurrence of the Shubnikov de-Hass (SdH) oscillation in the resistance at high fields. It should be noted that unlike for fields parallel to the layers, the resistive transition is relatively insensitive to the angles near $\theta $ = 180$^{\text{o}}$. The inset in Fig. 4 shows an expanded view of the resistive transitions at angles close to $\theta =$90$^{\text{o}}$ direction. With a slight increment in $\theta $, the transition is drastically broadened. The field component parallel to planes is almost constant for all angles shown in the inset and the maximum out of plane field component is about 0.5 T at $\theta $ = 91.50$^{\text{o}}$ and H = 30 T. {\it H}$_{c2}$ defined at the 1 $\Omega $ level as a function of angle is shown in Fig. 5 at a temperature of 1.56 K. Clearly, {\it H}$_{c2}$ decreases rapidly away from the parallel to the plane direction and is nearly saturated above 140$^{\text{o}}$. The three lines are fits to the three-dimensional anisotropic model (Eqn. (\ref{anisot})) and the decoupled layer results (Eqn. (\ref{tink}) and (\ref{pauli-tink})). The fits to Eqn. (% \ref{tink}) and (\ref{pauli-tink}) are indistinguishable in the scale shown. While all three fits seem reasonable at first sight, clear deviations are seen very close to $\theta =$90$^{\text{o}}$, as shown in the inset. A cusp-like feature is observed experimentally, as in the fit to the decoupled layer model, while the 3D fit is rounded with a negative curvature at the top. A better agreement with the data for the decoupled layer model at large angles is also evident with the 3D fit lying systematically under the data. At 1.56 K, the 2D fit gives $H_{c2\perp }$ = 2.27 T and $H_{c2\parallel }$ =24.5 T. Wanka et al.\cite{wanka} also found that the angular dependence for the X=I$_{3}$ salt was fit best by Eqn. (\ref{tink}). \section{Discussion} Our value of $H_{c2\perp }= 2.3$ T at 1.56 K can be compared with the value of about 1.8 T found from the irreversibility line deduced from torque measurements.\cite{sasaki} From the perpendicular upper critical field value of 2.3 T and Eqn. (\ref{hcperp}) we deduce an intralayer coherence length of 120 $\AA$. The anisotropic three-dimensional theory (Eqn. (\ref{ratio})) and the measured ratio of the upper critical fields gives a perpendicular coherence length of $\xi_\perp \simeq 13 \AA$. Since this is comparable to the interlayer spacing of $15 \AA$ we cannot expect the theory to apply. Hence, it is not surprising that the angular dependence is not described by Eqn. (% \ref{anisot}). If instead we consider the model of weakly coupled layers and use Eqn. (\ref {hcperp}) for the ratio of the critical fields we deduce that the thickness of the superconducting layer is $d = 40 \AA.$ Clearly, this is unrealistic because it should be smaller than the interlayer spacing. A more realistic value would be a few $\AA$. This suggests that the parallel upper critical field being determined by paramagnetic limiting rather than orbital effects is more realistic. Because of the extremely sensitive angular dependence of the resistive transition, a shoulder-like feature is developed in the resistive transition a few degrees away from the parallel to the plane direction. The upper critical field {\it H}$_{c2}$* can only be defined close to the planes. While the magnitude of {\it H}$_{c2}$* is larger than {\it H}$_{c2}$, as expected, it is difficult to distinguish the 2D and the 3D models with the available data. {\it H}$_{c2}$* decreases quasi-linearly with angle within the errors. The upper critical field determined from transport measurements has been under a lot of debate in the cuprate superconductors\cite{blumberg}. For field perpendicular to the planes, {\it H}$_{c2}(T)$ defined at certain fractional normal state resistance typically gives rise to a positive curvature at low temperatures. Various mechanisms have been proposed for the unconventional temperature dependence. However, it has been suggested that the {\it H}$_{c2}$ thus defined corresponds to the irreversibility or vortex melting line. For fields parallel to the layers, a vortex moving along the plane encounters negligible pinning as there is no normal core associated with Josephson vortices. Magnetization is practically always reversible in this orientation. The resistive onset field is clearly well separated from irreversibility field and reflects the true upper critical field. In the case of Sr$_{2}$RuO$_{4}$ and the quasi-one-dimensional organic superconductor (TMTSF)$_{2}$X where X=ClO$_{4}$ and PF$_{6}$\cite{lee}, the upper critical field in the plane has been found to exceed the Pauli limit, calculated from BCS theory. Combined with the strong dependence of the transition temperature on the impurity concentration and the temperature dependence of the Knight shift, triplet pairing or p-wave has been suggested in these systems. However, the quasi-linear temperature dependence observed here for both $H_{c2}$ and $H_{c2}$* is remarkably different from that of Sr$% _{2}$RuO$_{4}$ and Bechgaard salts. For both Sr$_{2}$RuO$_{4}$ and (TMTSF)$% _{2}$ClO$_{4}$, the $H_{c2}$ is found to saturate for $T/T_c < $ 0.2--0.4. While for (TMTSF)$_{2}$PF$_{6}$, $H_{c2}$(T) along both {\it a} and {\it b}' axes where X=ClO$_{4}$ displays a diverging temperature dependence near $T=$ 0 K. \section{conclusions} In summary, for fields parallel to the layers we have observed an upper critical field determined from resistive transition which is comparable to the paramagnetic limit estimated from thermodynamic quantities but is considerably larger than that calculated from BCS theory. There is no evidence of a first order transition in the field dependence of the resistivity, which would occur if there was a transition to a Fulde-Ferrell phase. The observed anisotropy of the upper critical field is much less than would be predicted by a model without paramagnetic limiting. The upper critical field determined is quasi-linear with temperature. The angular dependence of the resistive transition is consistent with the highly anisotropic nature of the title compound with a cusp-like angular dependence for field near the plane. \acknowledgements We thank P. Coleman, A. Dzurak, J. Merino, J. O'Brien and J. Wosnitza for helpful discussions. This work was supported in part by NSF grant No. DMR-9623306 and the Petroleum Research Fund ACS-PRF 33812-AC5. Work at the National High Magnetic Field Laboratory was supported by NSF Cooperative Agreement No. DMR-9016241 and the state of Florida. Work at UNSW was supported by the Australian Research Council. Work performed at Argonne National Laboratory was supported by the U.S. Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences, under Contract No. W-31-109-ENG-38.
\section{Introduction} In this paper we explore the possibility that the atmospheric and solar neutrino deficits can be produced by oscillations of the three known neutrinos with comparable mass splittings $\Delta m^2\sim 10^{-3}\,{\rm eV}^2$ and large $\nu_e/\nu_\mu$ and $\nu_\mu/\nu_\tau$ mixings. In section~\ref{sole} we recall why solar neutrino experiments do not exclude this possibility. In section~\ref{atm+sole} we show that this possibility is also perfectly consistent with atmospheric neutrino experiments, even if it is sometimes said that, since SuperKamiokande (SK) sees no anomaly in the rate of $\nu_e$ events, significant oscillations of atmospheric $\nu_e$ neutrinos are excluded. This is not the case~\cite{lungo}. In short the reason is the following: atmospheric neutrinos are produced by cosmic rays in the following proportion $$(N_{\nu_e},N_{\nu_\mu},N_{\nu_\tau})\propto(1,R,0).$$ Since $R\approx 2$ the nearly maximal $\nu_\mu/\nu_\tau$ oscillation responsible of the atmospheric $\nu$ anomaly gives oscillated neutrinos with composition $\propto(1,1,1)$ --- the only proportion not affected by further possible oscillations. Of course this argument is only approximate ($R$ is larger than 2 for $E_\nu\circa{>}1\,{\rm GeV}$ and the $\nu_\mu/\nu_\tau$ oscillation need not be exactly maximal): a numerical computation will confirm that the conclusion is correct. \smallskip From a theoretical point of view, the possibility of explaining neutrino anomalies with comparable $\Delta m^2$ has important consequences. It is easy to build models that naturally explain large mixings, but between neutrinos with comparable mass. It is also easy to build models that give hierarchical neutrinos, but with small mixings. It is more difficult to obtain large mixing angles ($\theta_{23}\sim 1$) between hierarchical neutrinos ($\Delta m^2_{23}\gg \Delta m^2_{12}$, all oscillation parameters are precisely defined later on in eq.\eq{V}): only few mass matrices (justifiable with various symmetries) naturally give this pattern~\cite{lungo,textures}. Thus, non hierarchical neutrinos would not give very restrictive indications on flavour physics. \medskip The paper is structured as follows. In section~\ref{sole} we explain why a large solar $\Delta m^2\sim10^{-3}\,{\rm eV}^2$ is not safely excluded by solar neutrino experiments. In section~\ref{fit} we describe how we will fit the SK data. Since this is a delicate task, in section~\ref{atm-fit} we test our procedure performing a complete fit of the most recent SK data, in the `standard' case where the solar $\Delta m^2_{12}$ is negligibly small. In section~\ref{atm+sole} we show that SK data are perfectly compatible with a $\Delta m^2_{12}$ as large as allowed by the {\sc Chooz} bound~\cite{CHOOZ}, and we discuss how its effects can be detected at SK and at future experiments. \section{Energy independent solar oscillations?}\label{sole} If solar and atmospheric neutrino experiments are not affected by unknown systematic errors or by rare statistical fluctuations, if the standard solar models (SSMs) are correct, if there are only the three known light neutrinos, then forthcoming neutrino experiments will confirm and measure more precisely that \begin{equation}\begin{array}{ll} \Delta m^2_{23}\approx 10^{-3}\,{\rm eV}^2 & \sin^2 2\theta_{23}\approx 1\\ \Delta m^2_{12}\approx 10^{-10}\,{\rm eV}^2 & \sin^2 2\theta_{12}\approx 1\\ \end{array}\end{equation} These values give a good fit of atmospheric neutrino data and give an acceptable fit ($10\%$ C.L.~\cite{BKS,noMSW,9903262}) of solar neutrino data. \medskip Since these conclusions are quite strong, it is useful to discuss if they are also strongly founded. \smallskip The well known `MSW solutions' with $\Delta m^2\approx 10^{-5}\,{\rm eV}^2$ give a poor fit of the distortion of the solar $^8$B spectrum observed by SuperKamiokande~\cite{noMSW,9903262} and are ruled out at 95\% C.L.~\cite{noMSW}. Maybe this experimental result (or the estimation of its uncertainties) is wrong. Maybe the distortion is not produced by $\nu$ oscillations, but is due to a flux of `hep' neutrinos $\sim15$ times higher than what predicted by SSMs~\cite{lungo,hep}. At the moment, it seems more safe to believe that there are at least three (instead of one) possible oscillation solutions to the solar neutrino problem. We now discuss why even this sentence is not strongly founded: it is not safely excluded that the solar neutrino anomaly can be explained by an energy independent $\nu_e\to \nu_e$ survival probability $P_{ee}\sim 1/2$ (as can be produced by a large $\Delta m^2_{12}\sim 10^{-3}\,{\rm eV}^2$). \smallskip The deficit $r_i=\Phi_i^{\rm exp}/\Phi_i^{\rm BP98}$ of solar neutrinos measured by the three kind of solar experiments ($i=$ Cl, Ga and SK), with respect to the central values $\Phi_i^{\rm BP98}$ of fluxes predicted by the BP98~\cite{BP98} SSM, are \begin{eqnsystem}{sys:sunexp} r_{\rm Cl}&=&0.315\pm0.025~~\cite{ClSun}\\ r_{\rm SK}&=&0.47\pm0.02~~~~~\cite{KaSun}\\ r_{\rm Ga}&=&0.58\pm0.05~~~~~\cite{GaSun} \end{eqnsystem} (the errors do not include the SSM uncertainty). The predictions of a solar-model-independent analysis, in presence of an energy-independent $P_{ee}$, are \begin{eqnsystem}{sys:sunth} r_{\rm Cl}&=& P_{ee}(0.03+0.73 R_{\rm {}^8B}+0.24 R_{\rm {}^7Be})\qquad\\ r_{\rm SK}&=& (0.15 +0.85 P_{ee}) R_{\rm {}^8B}\\ r_{\rm Ga}&=& P_{ee}(0.60+0.10 R_{\rm {}^7Be}+0.31 R_{\rm {}^8B}) \end{eqnsystem} where $R_\alpha\equiv \Phi_\alpha/\Phi_\alpha^{\rm BP98}$ is the ratio of total flux of type $\alpha$ neutrinos ($\alpha = \rm pp, p\hbox{$e$}p, ^7\!Be,^{13}\!N,^{15}\!O,{}^{17}\!F,{}^8\!B,\break h\hbox{$e$}p$) emitted by the sun, with respect to the central value of the BP98~\cite{BP98} SSM. Only two free parameters $ R_{\rm {}^8B}$ and $R_{\rm {}^7Be}$ appear in eq.s~(\ref{sys:sunth}); the others have been eliminated using the fact that the total luminosity of the sun is known, and other solid informations (see~\cite{CDFLR,lungo} for more details). \begin{figure*}[t] \begin{center} \begin{picture}(17.7,5) \putps(-0.5,0)(-0.5,0){fitchiq}{fitchiq.ps} \end{picture} \caption[SP]{\em $\chi^2$ of SuperKamiokande data as function of $\Delta m^2_{23}$ for $\theta_{23}=45\degree$, $\theta_{13}=0$ and negligible $\Delta m^2_{12}$. In fig.~{\rm\ref{fig:chiq}a (b)} we show the separate contribution from sub-GeV (multi-GeV) events; while all events are used in fig.~{\rm\ref{fig:chiq}c}. Continuous and dotted lines correspond to two different definitions of the $\chi^2$ (see the text). \label{fig:chiq}} \end{center}\end{figure*} \noindent From eq.s~(\ref{sys:sunexp}) and~(\ref{sys:sunth}) we can easily see that \begin{itemize} \item If BP98 is correct (it predicts $R_{\rm {}^8B}=1\pm 0.2$ and $R_{\rm {}^7Be}=1\pm0.05$~\cite{BP98}), the three experimental data are compatible with an energy-independent $P_{ee}$ only in presence of a very unprobable statistical fluctuation ($p\approx 0.2\%$)~\cite{PeeCteBad,lungo}; \item Even if SSMs are not correct (i.e.\ $R_{\rm {}^8B}$ and $R_{\rm {}^7Be}$ are treated as free parameters of order one), the experimental data are incompatible with an energy-independent $P_{ee}$~\cite{PeeCteBad,lungo}. A decent fit is possible only if SSMs are so wrong ($R_{\rm {}^7Be}$ close to zero) that the solar neutrino problem disappears~\cite{lungo} --- a possibility strongly disfavoured by recent helio-sysmological tests of SSMs. \item Solar data can be explained by an energy-inde\-pen\-dent $P_{ee}$ if BP98 is correct, but one of the experiments is affected by some unknown systematic error~\cite{lungo}. For example, {\em it is sufficient to double the error quoted by the chlorine experiment~\cite{ClSun} to have $P_{ee}=1/2$ compatible with experimental results} (in presence of a reasonably probable, $p\sim 10\%$, statistical fluctuation). \end{itemize} In conclusion we believe that the evidence for a deficit of solar neutrinos is strong, while there is yet no strong evidence for an energy dependent oscillation of solar neutrinos. \section{Fitting SK data}\label{fit} Reproducing what SK has really measured is a complex and delicate task. We briefly describe how we do the computation. Five basic ingredients are necessary for a fit of the SK data. \begin{enumerate} \item The experimental data: we use the most recent ones (736 days of data taking)~\cite{SKexp}. \item The prediction for the flux of atmospheric neutrinos produced by cosmic rays~\cite{Gaisser}. We include effects due to the magnetic field of the earth and to variation of solar activity. \item The oscillation probability for neutrinos across the earth and the atmosphere. In our case we have a generic neutrino $3\times 3$ mass matrix with 3 comparable $\Delta m^2$: since we know no simple analytic approximation that takes into account all potentially relevant matter effects (the MSW~\cite{MSW} effect and resonances that can affect the neutrinos that cross the mantle and the core of the earth~\cite{ParRes}) we include all matter effects with a fully numerical computation. The disadvantage is that the transition probabilities for oscillations with long pathlength $L\gg E_\nu /\Delta m^2$ are rapidly oscillating functions of the neutrino energy. Even for the simplest realistic model of earth density, a numerical averaging (using a sufficiently large number of $E_\nu$-bins) is more efficient than analytic averaging. \item The cross section and detection efficiencies in the SK detector. A simple and safe technique has been used in~\cite{SKfit} for fitting the old Kamio\-kande data. They employed the energy spectra of parent neutrinos given by the Monte Carlo simulation of the Kamiokande detector. The corresponding data for the much larger SK detector have not been published (without these information it is not possible to know what SK is really measuring), although they are now available at the www address~\cite{www}. We take into account that SK measures the neutrino direction with an error that depends on the $\nu$ energy~\cite{tesi} ($\delta \theta$ is around $60\degree$ in sub-GeV events and $17\degree$ in multi-GeV events). ~~We do not include in the fit data about `upward through going muons'~\cite{mu-roccie} because they are subject to larger theoretical uncertainties and they are too energetic for being strongly affected by oscillations (they are however very interesting for excluding alternative explanations of the atmospheric $\nu$ deficit~\cite{osc!}). \item A $\chi^2$ function. This is a delicate point, since it requires an estimation of theoretical uncertainties in neutrino fluxes and correlation between them. For simplicity we stick to the accurate definition of~\cite{fogli} and we will discuss when appropriate the effect of different definitions. \end{enumerate} Moreover we impose the {\sc Chooz} bound about disappearance of reactor $\bar{\nu}_e$~\cite{CHOOZ}. The {\sc Chooz} collaboration measures the annihilation energy $E=E_{{\bar \nu}_e}+m_{\rm p}-m_{\rm n}+(2-1)m_e$ of positrons produced by inverse beta decay $\bar{\nu}_e {\rm p}\to \bar{e}{\rm n}$ for various energy bins between $1$ and $7\,{\rm MeV}$. Since we will be interested in an oscillation pattern where the $\bar{\nu}_e$ survival probability depends on $E_{\bar{\nu}_e}$, we carefully treat the {\sc Chooz} data, grouping them into two $E_{\bar{\nu}_e}$ bins. We correctly reproduce the `initial' {\sc Chooz} bound ($\Delta m^2_{12}<0.9~10^{-3}\,{\rm eV}^2$ for maximal mixing at $90\%$ C.L.). A `final' analysis of the whole {\sc Chooz} data has not yet been presented: with increased statistics, the {\sc Chooz} bound could be improved up to $\approx0.6~10^{-3}\,{\rm eV}^2$~\cite{CHOOZ2}. \begin{figure*}[t] \begin{center} \begin{picture}(17.7,5) \putps(-0.5,0)(-0.5,0){fitstandard}{fitstandard.ps} \put(2.7,5.3){\ref{fig:fitstandard}a} \put(8.5,5.3){\ref{fig:fitstandard}b} \put(14.3,5.3){\ref{fig:fitstandard}c}\special{color cmyk 0 1. 1. 0.5} \put(14.3,4.5){excluded} \put(14.3,4){by {\sc Chooz}}\special{color cmyk 0 0 0 1.} \end{picture} \caption[SP]{\em Fit of SK and {\sc Chooz} data under the assumption that `solar' oscillations are negligible, so that the only relevant oscillations parameters are $\Delta m^2_{23}$, $\theta_{23}$ and $\theta_{13}$. Inside the darker (lighter) areas $\chi^2<24~(30)$ (the $\chi^2$ uses $22$ experimental data; the best fit has $\chi^2_{\rm min}=18$; In fig.~\hbox{\rm\ref{fig:fitstandard}a} the $\chi^2$ is minimized with respect to $\theta_{13}$, in fig.~\hbox{\rm\ref{fig:fitstandard}b} with respect to $\Delta m^2_{23}$, and in fig.~\hbox{\rm\ref{fig:fitstandard}c} with respect to $\theta_{23}$. \label{fig:fitstandard}} \end{center}\end{figure*} \section{Standard fit of SK data}\label{atm-fit} We now begin to study the SK data assuming, as usual, that $\Delta m^2_{12}$ is too small to have relevant effects. We start with a standard fit because, beyond being interesting, allows to check our computation with other ones~\cite{SKexp,fogli,otherSKfits}. We find a very satisfactory agreement with~\cite{fogli}, where many plots of oscillations effects are presented. The oscillation parameters are precisely defined in the following way. The neutrino mixing matrix $V$ is parametrized as \begin{equation} V = R_{23}(\theta_{23}) \pmatrix{1 &0&0 \cr 0&e^{i \phi}&0 \cr 0&0&1} R_{13}(\theta_{13}) R_{12}(\theta_{12}) \label{eq:V} \end{equation} where $R_{ij}(\theta_{ij})$ represents a rotation by $\theta_{ij}$ in the $ij$ plane and $i=\{1,2,3\}$ are three neutrino mass eigenstates of mass $m_i$. With this parameterization $\theta_{23}\sim45\degree$ gives the $\nu_\mu/\nu_\tau$ mixing tested at SK; while $\theta_{1i}$ produce the deficit of solar $\nu_e$ neutrinos ($\theta_{13}$ could be zero). We also define $\Delta m^2_{ij}\equiv m^2_j-m^2_i$. We can assume that $|\Delta m^2_{23}|$ is the largest splitting (see~\cite{lungo} for more details). With this parameterization atmospheric oscillations depend only on $\Delta m^2_{23}$, $\theta_{23}$ and $\theta_{13}$. \smallskip In fig.~1 the continuous lines show how the $\chi^2$ depends on $\Delta m^2_{23}$ (for maximal $\theta_{23}=45\degree$ and zero $\theta_{13}$) fitting separately the sub-GeV and the multi-GeV events. Compared to a similar fit done by the SK collaboration, our result is less sensitive to $\Delta m^2_{23}$ and allows slightly smaller values of $\Delta m^2_{23}$. This small difference is probably due to the fact that our fit does not include data about `upward through going muons'~\cite{mu-roccie}: since our $\chi^2$ has a particularly flat minimum these less significant data, that seem to prefer higher values of $\Delta m^2_{23}\sim 10^{-(3\div 1)}\,{\rm eV}^2$, can cause the small difference. The dotted lines in fig.~1 correspond to a simplified definition of the $\chi^2$: we fit separately the angular dependence of each one of the four kind of events measured by SK ($e$-like and $\mu$-like, sub-GeV and multi-GeV), treating the overall normalization of each one of them as free but including only statistical errors. We see that, at least in the fit in fig.~1, there is no significant difference between this simplified $\chi^2$ and the one in~\cite{fogli}. An interesting feature shown by fig.~\ref{fig:chiq} is that {\em small values of $\Delta m^2_{23}\circa{<}10^{-3}\,{\rm eV}^2$ (that are difficult to test at planned `long-baseline' experiments) are now disfavoured by the clean multi-GeV data only\/} (whose uncertainty is dominated by statistics), rather than by the sub-GeV data (whose interpretation strongly depends on the details of the SK experiment and on theoretical predictions about the neutrino flux). Thus the lower bound on $\Delta m^2$ is solid and can be improved with more statistics in the next years. \medskip In fig.~\ref{fig:fitstandard} we show the results of a fit in the relevant oscillations parameters, $\Delta m^2_{23}$, $\theta_{23}$ and $\theta_{13}$. The best fit has $\chi^2_{\rm min}=18$ (the $\chi^2$ uses 20 experimental data from SK and 2 from {\sc Chooz}). We show contour lines corresponding to the values $\chi^2=24$ and $\chi^2=30$. Using standard approximations, values of $\chi^2-\chi^2_{\rm min}$ can be converted into confidence levels that delimit `best fit regions', and values of $\chi^2$ can be converted into confidence levels that delimit `exclusion regions'. Shaded areas roughly correspond to $(90\div 99)\%$ confidence levels\footnote{We do not insist on the precise correspondence since it has no particular meaning. There is no objective way of converting $\chi^2$ levels into statements like ``oscillation parameters lie in the shaded region with $90\%$ probability''.}. In fig.~\hbox{\rm\ref{fig:fitstandard}a} we minimize the $\chi^2$ with respect to $\theta_{13}$ and determine the allowed regions in the plane $(\Delta m^2_{23},\theta_{23})$. We see that $\sin^2 2\theta_{23}>0.8$ at $90\%$ C.L. Fig.~\hbox{\rm\ref{fig:fitstandard}b} shows the $\chi^2$ minimized with respect to $\Delta m^2_{23}$ and determines the allowed values of the mixing angles. In fig.~\ref{fig:fitstandard}c we minimize the $\chi^2$ with respect to $\theta_{23}$ and show how the upper bound on $\theta_{13}$ depends on $\Delta m^2_{23}$. If $\Delta m^2_{23}=3~10^{-3}\,{\rm eV}^2$ the {\sc Chooz} bound (dashed lines) requires a small value of $\theta_{13}\circa{<}10\degree$. For a mass splitting below the {\sc Chooz} bound, $\Delta m^2_{23}\circa{<}10^{-3}\,{\rm eV}^2$, a larger $\theta_{13}$ is not forbidden by the {\sc Chooz} data, but disfavoured by SK because it generates an up/down asymmetry in the $e$-like (sub-GeV and multi-GeV) sample. However, as explained in the introduction, the sub-GeV asymmetry vanishes for $\theta_{23}=45\degree$, while the one in the multi-GeV sample vanishes for an appropriate value of $\theta_{23}\circa{<}45\degree$. Since our analysis does not say that $\Delta m^2_{23}=1\cdot 10^{-3}\,{\rm eV}^2$ is disfavoured, for an appropriate value of $\theta_{23}\approx 45\degree$, `large' $\theta_{13}\circa{>}20\degree$ are allowed. \begin{figure*}[t] \begin{center} \begin{picture}(17.7,5) \putps(-0.5,0)(-0.5,0){fit3masse}{fit3masse.ps} \put(2.7,5.3){$\ref{fig:fit3masse}a$} \put(8.5,5.3){$\ref{fig:fit3masse}b$} \put(14.3,5.3){$\ref{fig:fit3masse}c$}\special{color cmyk 0 1. 1. 0.5} \put(7.3,4.6){excluded by {\sc Chooz}} \put(13,4.6){excluded by {\sc Chooz}}\special{color cmyk 0 0 0 1.} \end{picture} \caption[SP]{\em Fit of SK data assuming that all $\Delta m^2$ are large enough to affect atmospheric neutrinos. As in fig.~\ref{fig:fitstandard} we show contour lines of the $\chi^2(\Delta m^2_{ij},\theta_{ij})$ at $\chi^2=24$ and $30$ ($\min\chi^2=17$). All fits are restricted to $\theta_{13}=0$ and $\theta_{12}=45\degree$. In fig.~\hbox{\rm\ref{fig:fit3masse}a} we show $\chi^2(\Delta m^2_{13},\theta_{23})$ for a large value of $\Delta m^2_{12}=0.8~10^{-3}\,{\rm eV}^2$; in fig.~\hbox{\rm\ref{fig:fit3masse}b (c)} we show the $\chi^2$ minimized with respect to $\Delta m^2_{13}$ ($\theta_{23}$). \label{fig:fit3masse}} \end{center}\end{figure*} \begin{figure*}[t] \begin{center} \begin{picture}(17.7,5.8) \putps(-0.2,0)(-0.2,0){f1245}{f1245.ps} \special{color cmyk 0 1. 1. 0.5} \put(2.4,5.8){$e$-sub GeV} \put(9.5,5.8){$e$-multi GeV}\special{color cmyk 1. 1. 0.2 0.2} \put(5.9,5.8){$\mu$-sub GeV} \put(13.1,5.8){$\mu$-multi GeV}\special{color cmyk 0 0 0 1.} \end{picture} \caption[SP]{\em Effect of `solar' oscillations on SK observables, normalized to the unoscillated rates, for $\theta_{12}=\theta_{23}=45\degree$, $\theta_{13}=0$, $\Delta m^2_{13}=3~10^{-3} \,{\rm eV}^2$ and $\Delta m^2_{12}=\{0,0.3,1,3,9\}10^{-4}\,{\rm eV}^2$ (larger values have longer dashing). The arrows on the horizontal axes denote the direction of incoming neutrinos. The crosses are the experimental data (their error bars only include statistical errors). \label{fig:sample}} \end{center} \begin{center} \begin{picture}(17.7,5.8) \putps(-0.2,0)(-0.2,0){f1230}{f1230.ps} \special{color cmyk 0 1. 1. 0.5} \put(2.4,5.8){$e$-sub GeV} \put(9.5,5.8){$e$-multi GeV}\special{color cmyk 1. 1. 0.2 0.2} \put(5.9,5.8){$\mu$-sub GeV} \put(13.1,5.8){$\mu$-multi GeV}\special{color cmyk 0 0 0 1.} \put(4.8,4.2){\footnotesize$9$} \put(4.8,3.95){\footnotesize$3$} \put(4.8,3.7){\footnotesize$1$} \put(4.75,3.45){\footnotesize$.3$} \put(2,3.2){\footnotesize$\Delta m^2_{12}/(10^{-4}\,{\rm eV}^2)=0$} \end{picture} \caption[SP]{\em As in fig.~\ref{fig:sample}, but for a smaller $\theta_{23}=30\degree$. The results for $\theta_{23}=60\degree$ are similar, but the sign of the effects is reversed. \label{fig:sample2}} \end{center}\end{figure*} \section{Non-standard fit of SK data}\label{atm+sole} Finally we assume that the `solar' $\Delta m^2_{12}$ is sufficiently large to affect atmospheric neutrinos\footnote{Related analyses, motivated by the fact that the large angle MSW solution allows $\Delta m^2_{12} < +0.2~10^{-3}\,{\rm eV}^2$ (the positive sign is the one that gives the desired MSW effect for solar neutrinos) have been recently performed in~\cite{Giunti,Smirnov}. For such values of $\Delta m^2_{12}$ our results agree with~\cite{Smirnov} and disagree with~\cite{Giunti}.}. The SK observables now depend on all oscillation parameters. For simplicity we only exhibit fits where $\theta_{12}=45\degree$ (as suggested by the deficit of solar neutrinos, if $\theta_{13}$ is small) and $\theta_{13}=0$, so that the CP-violating phase $\phi$ becomes irrelevant. In fig.s~\ref{fig:fit3masse} we show the result of our fit of SK atmospheric data in $\Delta m^2_{13}$, $\Delta m^2_{12}$ and $\theta_{23}$. Fig.~\ref{fig:fit3masse}a is done for $\Delta m^2_{12}=0.8~10^{-3}\,{\rm eV}^2$, just below the {\sc Chooz} bound, and shows that a good fit is possible for appropriate values of $\Delta m^2_{13}$ and $\theta_{23}$. In fig.~\ref{fig:fit3masse}b (c) we minimize the $\chi^2$ with respect to $\Delta m^2_{13}$ ($\theta_{23}$). The main result is that {\em values of $\Delta m^2_{12}$ as large as $10^{-3}\,{\rm eV}^2$ are compatible with the most recent SK data.} Infact the upper bound on $\Delta m^2_{12}\circa{<} 0.9~10^{-3}\,{\rm eV}^2$ --- shown in fig.s~\ref{fig:fit3masse}b,c --- comes from the {\sc Chooz} experiment. Even for the maximal value of $\Delta m^2_{12}$, `solar' oscillations have small effect on SK observables: they produce an up/down asymmetry of $e$-like events (this observable has a dominant statistical error) accompanied by a change in the overall number of events (this observable has a dominant theoretical error). Depending on how the $\chi^2$ is defined this small effect can slightly improve or deteriorate the fit. These small effects on $e$-like events produce the preference for values of $\theta_{23}\circa{<}50\degree$ when $\Delta m^2_{23}$ is larger, apparent from fig.s \ref{fig:fit3masse}a,b. We now discuss this up/down asymmetry of $e$-like events in more detail because it is the most promising signal of `solar' oscillations that can be observed at SK. \medskip The qualitative features of the asymmetry are well reproduced by the rough approximation (again obtained making the simplifying assumptions used in~\cite{lungo}) \begin{equation} {N_e^\uparrow\over N_e^\downarrow}\approx 1+\label{eq:Yesolar} \frac{R(1+\cos2\theta_{23})-2}{4}\sin^22\theta_{12}. \end{equation} As already explained in the introduction, if $R=2$ (as in the sub-GeV data) and $\theta_{23}=45\degree$ there is no effect. Multi-GeV neutrino events have $R\approx 3$; but are rarer and too energetic for being strongly affected by a $\Delta m^2_{12}$ below the {\sc Chooz} bound. Since the cancellation makes the situation intricate, a numerical computation is necessary to determine the possible signatures of a solar $\Delta m^2$ at SK. We show in figure~\ref{fig:sample} some example of how the quantities measured by SK are affected by `solar' oscillations for different values of $\Delta m^2_{12}$ if $\theta_{23}=45\degree$. The arrows on the horizontal axes denote the direction of the incoming neutrinos and correspond to the five bins of $\cos\vartheta_{\rm zenith}$ used by the SK collaboration to present their results (for example $\uparrow$ refers to the up-going neutrinos that cross the core of the earth). The rate of each one of the 20 bins is normalized with respect to the no oscillation case (our predictions for unoscillated rates are in satisfactory agreement with the Monte Carlo of the SK collaboration). We see that the effect is very small for any value of $\Delta m^2_{12}$. In figure~\ref{fig:sample2} we again show the effects of `solar' oscillations, but for the smallest value of $\theta_{23}=30\degree$ compatible with the SK data. There are now larger effects in the $e$ sub-GeV sample. We see that a small $\sin^22\theta_{23}=3/4$ gives a poorer fit of $\mu$ events; but this fit also depends on the value of $\Delta m^2_{23}$: for this reason we do not consider useful discussing small `solar' effects in the $\mu$ events. To correctly interpret fig.~\ref{fig:sample2}, we must remind that the overall normalization of the fluxes (i.e.\ the `1' line in the plot) has a $\sim20\%$ theoretical uncertainty. Moreover the `1' lines for the four different data samples can be moved independently by $\sim5\%$. Our $\chi^2$ knows that these systematic uncertainties are highly correlated and says that (with present statistics) even the `solar' effects shown in fig.~\ref{fig:sample2} are ``small effects''. For $\theta_{23}=60\degree$ the effects due to `solar' oscillations have similar size, but opposite sign. \medskip In all these computation we have assumed that $\Delta m_{23}^2,\Delta m^2_{12}>0$. Matter effects do not significantly affect the results if the $\Delta m^2$ have different signs. \begin{figure*}[t] \begin{center} \begin{picture}(17.7,5) \putps(-0.5,0)(-0.5,0){f1340}{f1340.ps} \putps(8.5,0)(8.5,0){f1350}{f1350.ps} \put(1.8,5.9){$\theta_{23}=40\degree$, $\theta_{13}=\{0,5\degree,10\degree,20\degree\}$} \put(10.5,5.9){$\theta_{23}=50\degree$, $\theta_{13}=\{0,5\degree,10\degree,20\degree\}$} \special{color cmyk 0 1. 1. 0.5} \put(0.5,1){$e$-subGeV} \put(9.5,1){$e$-subGeV} \put(4.1,1){$e$-multiGeV} \put(13.1,1){$e$-multiGeV}\special{color cmyk 1. 1. 0.2 0.2} \put(2.4,5.2){$\mu$-subGeV} \put(5.9,5.2){$\mu$-multiGeV} \put(11.3,5.2){$\mu$-subGeV} \put(14.9,5.2){$\mu$-multiGeV}\special{color cmyk 0 0 0 1.} \end{picture} \caption[SP]{\em Sample of how the SK observables are affected by a non zero $\theta_{13}$ for $\Delta m^2_{23}=3~10^{-3} \,{\rm eV}^2$ and $\theta_{23}=40\degree$ (left) $\theta_{23}=50\degree$ (right). The arrows on the horizontal axes denote the direction of the incoming neutrinos. \label{fig:sample13}} \end{center}\end{figure*} \subsection{Non zero $\theta_{13}$} So far we have assumed that $\theta_{13},\phi=0$. Like in the `standard' scenario, a small $\theta_{13}\circa{<}20\degree$ is allowed by the {\sc Chooz} and SK data. CP violation cannot affect SK observables since integration over neutrino energy averages it to zero. We now discuss how effects due to $\theta_{13}$ mixing can be distinguished from effects due to a `solar' oscillation. In fig.~\ref{fig:sample13} we show how a $\theta_{13}=\{0,5\degree,10\degree,20\degree\}$ affects the SK observables if $\Delta m^2_{12}\ll\Delta m^2_{23}=3\*10^{-3}\,{\rm eV}^2$, $\theta_{23}=\{40\degree,50\degree\}$. Comparing the results of the numerical computation shown in fig.~\ref{fig:sample13} with fig.s~\ref{fig:sample},\ref{fig:sample2}, we notice two main differences \begin{itemize} \item $\theta_{13}$ oscillations mainly affect multi-GeV $e$-like events, while `solar' oscillations can only affect $e$ sub-GeV events. \item For a given value of $\theta_{23}$, `solar oscillations' and `$\theta_{13}$-oscillations' produce up/down asymmetries of opposite sign. The up/down asymmetry produced by $\theta_{13}$ is different from zero and positive even if $\theta_{23}=45\degree$. \end{itemize} (In particular, a not too large $\theta_{13}$ can only produce a few $\%$ excess of sub-GeV $e$-like events). Both these features can be understood comparing the approximate up/down asymmetries produced by a small $\theta_{13}$, \begin{equation} {N_e^\uparrow\over N_e^\downarrow}\approx 1+2\theta_{13}^2(R\sin^2\theta_{23}-1),\label{eq:Ye13} \end{equation} with the corresponding approximation for `solar' effects, eq.\eq{Yesolar}. We remind that $\Delta m^2_{23}\circa{>} \Delta m^2_{12}$ and that sub-GeV events have $R\approx 2$, while multi-GeV ones have $R\approx 3$. In both cases the effects cancel out if $R\approx 2$ and $\theta_{23}\approx 45\degree$. \section{Conclusions} In conclusion a large `solar' $\Delta m^2_{12}\circa{<} 10^{-3}\,{\rm eV}^2$ is not safely excluded by solar neutrino experiments and is allowed by atmospheric neutrino experiments\footnote{On the contrary no acceptable fit of SK data is possible for the mass pattern $\Delta m^2_{12}\sim10^{-3}\,{\rm eV}^2$ and $\Delta m^2_{23}\sim \,{\rm eV}^2$, sometimes invoked for reconciling the unconfirmed LSND oscillation with solar and atmospheric ones. We do not show any numerical result because this fact is sufficiently clear from the approximate analysis in~\cite{lungo}.} A `solar' oscillation has little effect on SK observables and can slightly ameliorate or deteriorate the fit of SK data, depending on how the $\chi^2$ is defined. Its most clear signature at SK is an up to $\sim15\%$ angular dependent excess (or deficit) of $e$ sub-GeV events. An indication for a $\sim10\%$ excess of $e$-like events in the sub-GeV sample was present in the first year of data taking at SK; but the evidence has decreased in the most recent analyses with doubled statistics. However this happened in a not very nice way: \begin{itemize} \item The SK collaboration has introduced small improvements in their Monte carlo, and obtained slightly different predictions; \item The rate of $e$ sub-GeV events in the most recent part of the sample (last 321 days of data taking) is $(18\pm 5)\%$ lower than in the first part (first 414 days)~\cite{LoSecco}. \end{itemize} Moreover, if the overall flux of atmospheric $\nu$ is somewhat lower than what current estimates indicate --- as suggested by a recent (preliminary) measurement of cosmic ray fluxes~\cite{BESS} --- than SK is observing a smaller deficit of $\nu_\mu$ and an excess of $\nu_e$ events. In conclusion we believe that the normalization of $\nu$ fluxes is still very uncertain and that an excess (or even a deficit) of $e$ sub-GeV events could be present. With more statistics it will be possible to use the unoscillated down-going multi-GeV events to fix the normalization, or to search for up/down asymmetries in the $e$ samples. \medskip To conclude, a large `solar' $\Delta m^2_{12}$ can be experimentally investigated in different ways: \begin{itemize} \item {\bf SuperKamiokande}, when high-statistics will be available, could see some indication in the details of $e$ events; \item future {\bf solar experiments} (like Borexino) could exclude this possibility; \item {\bf KamLand} could soon observe an evident deficit of reactor $\bar{\nu}_e$; \item `{\bf long-baseline experiments}' (like {\sc K2K} and {\sc Minos}) can confirm the $\nu_\mu \leftrightarrow\nu_\tau$ oscillation seen at SK. These experiments could also see the $\nu_e$ appearance signal due to a sufficiently large $\theta_{13}$. On the contrary $\nu_e$ appearance\footnote{The $\nu_\mu$ disappearence observed at SK reduces the $\nu_e$ appearance signals at {\sc K2K} and {\sc Minos} by a factor close to $2$. This explains why, according to fig~\ref{fig:fMinos}b, these experiments are less sensitive to $\nu_\mu\to\nu_e$ oscillations than in standard estimates, done using the `two-neutrino approximation' where $\nu_\tau$ are neglected.} due only to `solar' effects can be seen only if $\Delta m^2_{12}$ is very close to its maximal value, $10^{-3}\,{\rm eV}^2$. CP violating effects in $\nu_\mu\to \nu_e$ are too small to affect the experiment, except maybe in some extreme case~\cite{rom}. All these conclusions are illustrated by fig.~\ref{fig:fMinos}, where we show contour-plots of the relevant transition probabilities averaged over the energy spectrum of the `medium energy' $\nu_\mu$ beam at {\sc Minos} (see the caption for more details). \item future `{\bf neutrino factories}' (beam of $\bar{\nu}_\mu\nu_e$ produced by decay of muons)~\cite{nu-factory}, if $\Delta m^2_{12}$ is large enough, could study CP violation and make precision measurements of the oscillation parameters (but it could be difficult to distinguish $\theta_{13}$ effects from $\Delta m^2_{12}$ effects). \end{itemize} \begin{figure*}[t] \begin{center} \begin{picture}(17.7,4) \putps(-0.5,0)(-0.5,0){fMinos}{fMinos.ps} \put(2,5.3){\ref{fig:fMinos}a: $\nu_\mu\to\nu_\tau$} \put(7.7,5.3){\ref{fig:fMinos}b: $\nu_\mu\to\nu_e$} \put(13.1,5.3){\ref{fig:fMinos}c: CP in $\nu_\mu\to\nu_e$}\special{color cmyk 0 1. 1. 0.5} \special{color cmyk 0 0 0 1.} \end{picture} \caption[SP]{\em Contour plots of the transition probabilities $P(\nu_\mu\to \nu_\tau)$ \hbox{(fig.~\rm\ref{fig:fMinos}a)}, CP-conserving part of $P(\nu_\mu\to \nu_e)$ \hbox{(fig.~\rm\ref{fig:fMinos}b)} and CP-violating part of $P(\nu_\mu\to \nu_e)$ \hbox{(fig.~\rm\ref{fig:fMinos}c)} at {\sc Minos} (`medium' beam) as function of the mass splittings $(\Delta m^2_{12},\Delta m^2_{23})$ for maximal CP-violatng phase $\phi=90\degree$, maximal $\theta_{12}=\theta_{23}=\pi/4$ and moderately large $\theta_{13}=20\degree$. For this value of $\theta_{13}$, $\Delta m^2_{23}\circa{>}3~10^{-3}$ is disfavoured by {\sc Chooz}. {\sc Minos} can explore the shaded regions. \label{fig:fMinos}} \end{center}\end{figure*} \paragraph{Note added} In the recent paper~\cite{3maxmix} it was observed that if $\Delta m^2_{23}\approx 10^{-3}\,{\rm eV}^2$ the (interesting?) scenario named `tri-maximal mixing' (in our language $\theta_{12}=\theta_{23}=\pi/4$ and $\sin^2\theta_{13}=1/3$, i.e.\ $\theta_{13}\approx 35\degree$) is still compatible with SK and {\sc Chooz} data (at least until the complete {\sc Chooz} data will be presented), due to certain cancellations. This is also shown by fig.~\ref{fig:fitstandard}c: as we have discussed, these cancellations are not a special feature of tri-maximal mixing. Quite generally SK is rather insensitive to new oscillations beyond the observed one. \paragraph{Acknowledgments} I thank R. Barbieri, D. Nicol\`o and A. Yu Smirnov for clarifying discussions. \frenchspacing \small\footnotesize
\section{Motivation and summary of results}\label{sec:intro} A general goal in the design of randomized algorithms is to obtain fast algorithms with {\em small error probabilities}. Along these lines is also the goal of obtaining fast algorithms that are zero-error (a.k.a.~Las Vegas), as opposed to bounded-error (a.k.a.~Monte Carlo). We examine these themes in the context of {\em quantum\/} algorithms, and present a number of new upper and lower bounds that contrast with those that arise in the classical case. The error probabilities of many classical probabilistic algorithms can be reduced by techniques that are commonly referred to as {\em amplification}. For example, if an algorithm $A$ that errs with probability $\le {1 \over 3}$ is known, then an error probability bounded above by an arbitrarily small $\varepsilon>0$ can be obtained by running $A$ independently $\Theta(\log({1/\varepsilon}))$ times and taking the majority value of the outcomes. This amplification procedure increases the running time of the algorithm by a multiplicative factor of $\Theta(\log({1/\varepsilon}))$ and is optimal (assuming that $A$ is only used as a black-box). We first consider the question of whether or not it is possible to perform amplification more efficiently on a {\em quantum\/} computer. A classical probabilistic algorithm $A$ is said to {\em $(p,q)$-compute} a function $f : \01^{\ast} \rightarrow \01$ if \begin{eqnarray*} \label{classical} \Pr[A(x)=1] \!& &\! \cases{ \,\le p & if $f(x) = 0$\cr \,\ge q & if $f(x) = 1$.\cr} \end{eqnarray*} Algorithm $A$ can be regarded as a {\em deterministic} algorithm with an auxiliary input $r$, which is uniformly distributed over some underlying sample space $S$ (usually $S$ is of the form $\01^{l(|x|)}$). We will focus our attention on the {\em one-sided-error\/} case (i.e.\ when $p=0$) and prove bounds on quantum amplification by translating them to bounds on quantum search. In this case, for any $x\in\01^n$, $f(x)=1$ iff $(\exists r \in S)(A(x,r)=1)$. Grover's quantum search algorithm~\cite{grover:search} (and some refinements of it~\cite{bbht:bounds,brassard&hoyer:simon,bht:counting,mosca:eigen, grover:framework}) can be cast as a quantum amplification method that is provably more efficient than any classical method. It amplifies a $(0,q)$-algorithm to a $(0,\textstyle{1 \over 2})$-quantum-computer with $O(1/\sqrt{q})$ executions of $A$, whereas classically $\Theta(1/q)$ executions of $A$ would be required to achieve this. It is natural to consider other amplification problems, such as amplifying $(0,q)$-computers to $(0,1 - \varepsilon)$-quantum-computers ($0<q<1 - \varepsilon<1$). We give a tight analysis of this. \begin{theorem}\label{amp} Let $A : \01^n \times S \rightarrow \{0,1\}$ be a classical probabilistic algorithm that $(0,q)$-computes some function $f$, and let $N=|S|$ and $\varepsilon\geq 2^{-N}$. Then, given a black-box for $A$, the number of calls to $A$ that are necessary and sufficient to $(0,1-\varepsilon)$-quantum-compute $f$ is \begin{eqnarray}\label{ampbound} \Theta\left(\sqrt{N}\left({\textstyle{\sqrt{\log(1 / \varepsilon) + q N}}} - \sqrt{q N}\right)\right). \end{eqnarray} \end{theorem} The lower bound is proven via the polynomial method~\cite{nisan&szegedy:degree,bbcmw:polynomials} and with adaptations of techniques from~\cite{paturi:degree,coppersmith&rivlin:poly}. The upper bound is obtained by a combination of ideas, including repeated calls to an exact quantum search algorithm for the special case where the exact number of solutions is known~\cite{brassard&hoyer:simon,bht:counting}. {}From Theorem~\ref{amp} we deduce that amplifying $(0,\textstyle{1 \over 2})$ classical computers to $(0,1-\varepsilon)$ quantum computers requires $\Theta(\log(1/\varepsilon))$ executions, and hence cannot be done more efficiently in the quantum case than in the classical case. These bounds also imply a remarkable algorithm for amplifying a classical $(0,{1 \over N})$-computer $A$ to a $(0,1-\varepsilon)$ quantum computer. Note that if we follow the natural approach of composing an optimal $(0,{1\over N}) \rightarrow (0,\textstyle{1 \over 2})$ amplifier with an optimal $(0,\textstyle{1 \over 2}) \rightarrow (0,1-\varepsilon)$ amplifier then our amplifier makes $\Theta(\sqrt{N}\log(1 / \varepsilon))$ calls to $A$. On the other hand, Theorem~1 shows that, in the case where $N = |S|$, there is a more efficient $(0,{1 \over N}) \rightarrow (0,1-\varepsilon)$ amplifier that makes only $\Theta(\sqrt{N\log(1/\varepsilon)})$ calls to $A$ (and this is optimal). Next we turn our attention to the {\em zero-error\/} (Las Vegas) model. A zero-error algorithm never outputs an incorrect answer but it may claim ignorance (output `inconclusive') with probability $\leq 1/2$. Suppose we want to compute some function $f:\01^N\rightarrow\01$. The input $x\in\01^N$ can only be accessed by means of queries to a black-box which returns the $i$th bit of $x$ when queried on $i$. Let $D(f)$ denote the number of variables that a {\em deterministic} classical algorithm needs to query (in the worst case) in order to compute $f$, $R_0(f)$ the number of queries for a {\em zero-error} classical algorithm, and $R_2(f)$ for {\em bounded-error}. There is a monotone function $g$ with $R_0(g)\in O(D(g)^{0.753\ldots})$~\cite{snir:trees,saks&wigderson:trees}, and it is known that $R_0(f) \geq \sqrt{D(f)}$ for any function $f$~\cite{blum&impagliazzo:generic,hh:oneway}. It is a longstanding open question whether $R_0(f)\geq\sqrt{D(f)}$ is tight. We solve the analogous question for monotone functions for the quantum case. Let $Q_E(f)$, $Q_0(f)$, $Q_2(f)$ respectively be the number of queries that an exact, zero-error, or bounded-error quantum algorithm must make to compute $f$. For zero-error quantum algorithms, there is an issue about the precision with which its gates are implemented: any slight imprecisions can reduce an implementation of a zero-error algorithm to a bounded-error one. We address this issue by requiring our zero-error quantum algorithms to be {\em self-certifying\/} in the sense that they produce, with constant probability, a {\em certificate\/} for the value of $f$ that can be verified by a {\em classical\/} algorithm. As a result, the algorithms remain zero-error even with imperfect quantum gates. The number of queries is then counted as the {\em sum} of those of the quantum algorithm (that searches for a certificate) and the classical algorithm (that verifies a certificate). Our upper bounds for $Q_0(f)$ will all be with self-certifying algorithms. We first show that $Q_0(f)\geq \sqrt{D(f)}$ for every {\em monotone} $f$ (even without the self-certifying requirement). Then we exhibit a family of monotone functions that nearly achieves this gap: for every $\varepsilon>0$ we construct a $g$ such that $Q_0(g)\in O(D(g)^{0.5 + \varepsilon})$. In fact even $Q_0(g)\in O(R_2(g)^{0.5 + \varepsilon})$. These $g$ are so-called ``AND-OR-trees''. They are the first examples of functions $f:\01^N \rightarrow \01$ whose quantum zero-error query complexity is asymptotically less than their classical zero-error or bounded-error query complexity. It should be noted that $Q_0(\mbox{OR})=N$~\cite{bbcmw:polynomials}, so the quadratic speedup from Grover's algorithm is lost when zero-error performance is required. Furthermore, we apply the idea behind the above zero-error quantum algorithms to obtain a new result in communication complexity. We derive from the AND-OR-trees a communication complexity problem where an asymptotic gap occurs between the zero-error quantum communication complexity and the zero-error classical communication complexity (there was a previous example of a zero-error gap for a function with restricted domain in~\cite{BuhrmanCleveWigderson98} and bounded-error gaps in~\cite{astu:qsampling,raz:qcc}). This result includes a new lower bound in classical communication complexity. We also state a result by Hartmut Klauck, inspired by an earlier version of this paper, which gives the first {\em total} function with quantum-classical gap in the zero-error model of communication complexity. Finally, a class of black-box problems that has received wide attention concerns the determination of monotone graph properties \cite{rivest&vuillemin,kss:topo,king:graphprop,hajnal:graphprop}. Consider a directed graph on $n$ vertices. It has $n(n-1)$ possible edges and hence can be represented by a black-box of $n(n-1)$ binary variables, where each variable indicates whether or not a specific edge is present. A {\em nontrivial monotone graph property\/} is a property of such a graph (i.e.~a function $P : \{0,1\}^{n(n-1)} \rightarrow \{0,1\}$) that is non-constant, invariant under permutations of the vertices of the graph, and monotone. Clearly, $n(n-1)$ is an upper bound on the number of queries required to compute such properties. The {\em Aanderaa-Karp-Rosenberg} or {\em evasiveness conjecture} states that $D(P)=n(n-1)$ for all $P$. The best known general lower bound is $\Omega(n^2)$~\cite{rivest&vuillemin,kss:topo,king:graphprop}. It has also been conjectured that $R_0(P)\in\Omega(n^2)$ for all $P$, but the current best bound is only $\Omega(n^{4/3})$~\cite{hajnal:graphprop}. A natural question is whether or not quantum algorithms can determine monotone graph properties more efficiently. We show that they can. Firstly, in the exact model we exhibit a $P$ with $Q_E(P)<n(n-1)$, so the evasiveness conjecture fails in the case of quantum computers. However, we also prove $Q_E(P)\in\Omega(n^2)$ for all $P$, so evasiveness does hold up to a constant factor for exact quantum computers. Secondly, we give a nontrivial monotone graph property for which the evasiveness conjecture is violated by a zero-error quantum algorithm: let STAR be the property that the graph has a vertex which is adjacent to all other vertices. Any classical (zero-error or bounded-error) algorithm for STAR requires $\Omega(n^2)$ queries. We give a zero-error quantum algorithm that determines STAR with only $O(n^{3/2})$ queries. Finally, for bounded-error quantum algorithms, the OR problem trivially translates into the monotone graph property ``there is at least one edge'', which can be determined with only $O(n)$ queries via Grover's algorithm~\cite{grover:search}. \section{Basic definitions and terminology} See~\cite{berthiaume:qc,bbcmw:polynomials} for details and references for the quantum circuit model. For $b\in\01$, a query gate $O$ for an input $x=(x_0,\ldots,x_{N-1})\in\01^N$ performs the following mapping, which is our only way to access the bits $x_j$: $$ \st{j,b}\rightarrow\st{j,b\oplus x_j}. $$ We sometimes use the term ``black-box'' for $x$ as well as $O$. A quantum algorithm or gate network $A$ with $T$ queries is a unitary transformation $A=U_TOU_{T-1}O\ldots OU_1OU_0$. Here the $U_i$ are unitary transformations that do not depend on $x$. Without loss of generality we fix the initial state to $\st{\vec{0}}$, independent of $x$. The final state is then a superposition $A\st{\vec{0}}$ which depends on $x$ only via the $T$ query gates. One specific qubit of the final state (the rightmost one, say) is designated for the output. The acceptance probability of a quantum network on a specific black-box $x$ is defined to be the probability that the output qubit is 1 (if a measurement is performed on the final state). We want to compute a function $f:\01^N\rightarrow\01$, using as few queries as possible (on the worst-case input). We distinguish between three different error-models. In the case of {\em exact} computation, an algorithm must always give the correct answer $f(x)$ for every $x$. In the case of {\em bounded-error} computation, an algorithm must give the correct answer $f(x)$ with probability $\geq 2/3$ for every $x$. In the case of {\em zero-error} computation, an algorithm is allowed to give the answer `don't know' with probability $\leq 1/2$, but {\em if} it outputs an answer (0 or 1), then this must be the correct answer. The complexity in this zero-error model is equal up to a factor of 2 to the {\em expected} complexity of an optimal algorithm that always outputs the correct answer. Let $D(f)$, $R_0(f)$, and $R_2(f)$ denote the exact, zero-error and bounded-error classical complexities, respectively, and $Q_E(f)$, $Q_0(f)$, $Q_2(f)$ be the corresponding quantum complexities. Note that $N\geq D(f)\geq Q_E(f)\geq Q_0(f)\geq Q_2(f)$ and $N\geq D(f)\geq R_0(f)\geq R_2(f)\geq Q_2(f)$ for every $f$. \section{Tight trade-offs for quantum searching}\label{secerrorbound} In this section, we prove Theorem~\ref{amp}, stated in Section~\ref{sec:intro}. The {\em search} problem is the following: for a given black-box $x$, find a $j$ such that $x_j=1$ using as few queries to $x$ as possible. A quantum computer can achieve error probability $\leq 1/3$ using $T\in\Theta(\sqrt{N})$ queries~\cite{grover:search}. We address the question of how large the number of queries should be in order to be able to achieve a very small error $\varepsilon$. We will prove that if $T<N$, then $ T\in\Theta\left(\sqrt{N\log(1/\varepsilon)}\right). $ This result will actually be a special case of a more general theorem that involves a promise on the number of solutions. Suppose we want to search a space of $N$ items with error $\varepsilon$, and we are promised that there are at least some number $t<N$ solutions. The higher $t$ is, the fewer queries we will need. In the appendix we give the following lower bound on $\varepsilon$ in terms of $T$, using tools from~\cite{bbcmw:polynomials,paturi:degree,coppersmith&rivlin:poly}. \begin{theorem}\label{thqerrorboundt} Under the promise that the number of solutions is at least $t$, every quantum search algorithm that uses $T\leq N-t$ queries has error probability $$ \varepsilon\in\Omega\left(e^{-4bT^2/(N-t)-8T\sqrt{tN/(N-t)^2}}\right). $$ \end{theorem} Here $b$ is a positive universal constant. This theorem implies a lower bound on $T$ in terms of $\varepsilon$. To give a tight characterization of the relations between $T$, $N$, $t$ and $\varepsilon$, we need the following upper bound on $T$ for the case $t=1$: \begin{theorem}\label{thupperboundT} For every $\varepsilon>0$ there exists a quantum search algorithm with error probability $\leq\varepsilon$ and $\displaystyle O\left(\sqrt{N\log(1/\varepsilon)}\right)$ queries. \end{theorem} \begin{proof} Set $t_0=\ceil{\log(1/\varepsilon)}$. Consider the following algorithm: \begin{enumerate} \item Apply exact search for $t=1,\ldots,t_0$, each of which takes $O(\sqrt{N/t})$ queries. \item If no solution has been found, then conduct $t_0$ searches, each with $O(\sqrt{N/t_0})$ queries. \item Output a solution if one has been found, otherwise output `no'. \end{enumerate} The query complexity of this algorithm is bounded by $$ \sum_{t=1}^{t_0}O\left(\sqrt{\frac{N}{t}}\right)+ t_0 O\left(\sqrt{\frac{N}{t_0}}\right) =O\left(\sqrt{N\log(1/\varepsilon)}\right). $$ If the real number of solutions was in $\{1,\ldots,t_0\}$, then a solution will be found with certainty in step~1. If the real number of solutions was $>t_0$, then each of the searches in step~2 can be made to have error probability $\leq 1/2$, so we have total error probability at most $(1/2)^{t_0}\leq\varepsilon$. \end{proof} A more precise analysis gives $T\leq 2.45\sqrt{N\log(1/\varepsilon)}$. It is interesting that we can use this to prove something about the constant $b$ of the Coppersmith-Rivlin theorem (see appendix): for $t=1$ and $\varepsilon\in o(1)$, the lower bound asymptotically becomes $T\geq\sqrt{N\log(1/\varepsilon)/4b}$. Together these two bounds imply $b\geq 1/4(2.45)^2\approx 0.042$. The main theorem of this section tightly characterizes the various trade-offs between the size of the search space $N$, the promise $t$, the error probability $\varepsilon$, and the required number of queries: \begin{theorem}\label{thm:tight} Fix $\eta\in(0,1)$, and let $N>0$, $\varepsilon\geq 2^{-N}$, and $t\leq\eta N$. Let $T$ be the optimal number of queries a quantum computer needs to search with error $\leq\varepsilon$ through an unordered list of $N$ items containing at least $t$ solutions. Then $$ \log(1/\varepsilon)\in\Theta\left(\frac{T^2}{N}+T\sqrt{\frac{t}{N}}\right). $$ \end{theorem} \begin{proof} {}From Theorem~\ref{thqerrorboundt} we obtain the upper bound $\displaystyle \log(1/\varepsilon)\in O\left(\frac{T^2}{N}+T\sqrt{\frac{t}{N}}\right).$ To prove a lower bound on $\log(1/\varepsilon)$ we distinguish two cases. {\bf Case 1: $T\geq\sqrt{tN}$.} By Theorem~\ref{thupperboundT}, we can achieve error $\leq\varepsilon$ using $T_u\in O(\sqrt{N\log(1/\varepsilon)})$ queries. Now (leaving out some constant factors): $$ \log(1/\varepsilon)\geq\frac{T_u^2}{N}\geq \frac{1}{2}\left(\frac{T^2}{N}+T\frac{T}{N}\right)\geq \frac{1}{2}\left(\frac{T^2}{N}+T\sqrt{\frac{t}{N}}\right). $$ {\bf Case 2: $T<\sqrt{tN}$.} We can achieve error $\leq 1/2$ using $O(\sqrt{N/t})$ queries, and then classically amplify this to error $\leq 1/\varepsilon$ using $O(\log(1/\varepsilon))$ repetitions. This takes $T_u\in O(\sqrt{N/t}\log(1/\varepsilon))$ queries in total. Now: $$ \log(1/\varepsilon)\geq T_u\sqrt{\frac{t}{N}}\geq \frac{1}{2}\left(T\sqrt{\frac{t}{N}}+T\sqrt{\frac{t}{N}}\right)\geq $$ $$ \frac{1}{2}\left(\frac{T^2}{N}+T\sqrt{\frac{t}{N}}\right). $$ \vspace*{-7mm} \end{proof} Rewriting Theorem~\ref{thm:tight} (with $q=t/N$) yields the general bound of Theorem~\ref{amp}. For $t=1$ this becomes $T\in\Theta(\sqrt{N\log(1/\varepsilon)})$. Thus no quantum search algorithm with $O(\sqrt{N})$ queries has error probability $o(1)$. Also, a quantum search algorithm with $\varepsilon\leq 2^{-N}$ needs $\Omega(N)$ queries. For the case $\varepsilon=1/3$ we re-derive the bound $\Theta(\sqrt{N/t})$ from~\cite{bbht:bounds}. \section{Applications of Theorem 1 to amplification}\label{secnoampl} In this section we apply the bounds from Theorem~\ref{amp} to examine the speedup possible for amplifying classical one-sided error algorithms via quantum algorithms. Observe that searching for items in a search space of size $N$ and figuring out whether a probabilistic one-sided error algorithm $A$ with sample space $S$ of size $N$ accepts are essentially the same thing. Let us analyze some special cases more closely. Suppose that we want to amplify an algorithm $A$ that $(0,\textstyle{1 \over 2})$-computes some function $f$ to $(0,1-\varepsilon)$. Then substituting $|S| = N$ and $q = \textstyle{1 \over 2}$ into Eq.~(\ref{ampbound}) in Theorem~\ref{amp} yields \begin{theorem}\label{thnoamplrp} Let $A : \01^n \times S \rightarrow \{0,1\}$ be a classical probabilistic algorithm that $(0,\textstyle{1 \over 2})$-computes some function $f$, and $\varepsilon\geq 2^{-|S|}$. Then, given a black-box for $A$, the number of calls to $A$ that any quantum algorithm needs to make to $(0,1-\varepsilon)$-compute $f$ is $\Omega(\log(1/\varepsilon))$. \end{theorem} Hence amplification of one-sided error algorithms with fixed initial success probability cannot be done more efficiently in the quantum case than in the classical case. Since one-sided error algorithms are a special case of bounded-error algorithms, the same lower bound also holds for amplification of bounded-error algorithms. A similar but slightly more elaborate argument as above shows that a quantum computer still needs $\Omega(\log(1/\varepsilon))$ applications of $A$ when $A$ is zero-error. Some other special cases of Theorem~\ref{amp}: in order to amplify a $(0,{1 \over N})$-computer $A$ to a $(0,\textstyle{1 \over 2})$-computer, $\Theta(\sqrt{N})$ calls to $A$ are necessary and sufficient (and this is essentially a restatement of known results of Grover and others about quantum searching~\cite{grover:search,bbht:bounds}). Also, in order to amplify a $(0,{1 \over N})$-computer with sample space of size $N$ to a $(0,1 - \varepsilon)$-computer, $\Theta(\sqrt{N \log(1 / \varepsilon)})$ calls to $A$ are necessary and sufficient. Finally, consider what happens if the size of the sample space is unknown and we only know that $A$ is a classical one-sided error algorithm with success probability $q$. Quantum amplitude amplification can improve the success probability to $1/2$ using $O(1/\sqrt{q})$ repetitions of $A$. We can then classically amplify the success probability further to $1-\varepsilon$ using $O(\log(1/\varepsilon))$ repetitions. In all, this method uses $O(\log(1/\varepsilon)/\sqrt{q})$ applications of $A$. Theorem~\ref{thm:tight} implies that this is best possible in the worst case (i.e.~if $A$ happens to be a classical algorithm with very large sample space). \section{Zero-error quantum algorithms}\label{secsepzero} In this section we consider zero-error complexity of functions in the query (a.k.a. black-box) setting. The best general bound that we can prove between the quantum zero-error complexity $Q_0(f)$ and the classical deterministic complexity $D(f)$ for total functions is the following (the proof is similar to the $D(f)\in O(Q_E(f)^4)$ result given in~\cite{bbcmw:polynomials} and uses an unpublished proof technique of Nisan and Smolensky): \begin{theorem} For every total function $f$ we have $D(f)\in O(Q_0(f)^4)$. \end{theorem} We will in particular look at {\em monotone increasing} $f$. Here the value of $f$ cannot flip from 1 to 0 if more variables are set to 1. For such $f$, we improve the bound to: \begin{theorem}\label{thdvsqo} For every total monotone Boolean function $f$ we have $D(f)\leq Q_0(f)^2$. \end{theorem} \begin{proof} Let $s(f)$ be the {\em sensitivity} of $f$: the maximum, over all $x$, of the number of variables that we can individually flip in $x$ to change $f(x)$. Let $x$ be an input on which the sensitivity of $f$ equals $s(f)$. Assume without loss of generality that $f(x)=0$. All sensitive variables must be 0 in $x$, and setting one or more of them to 1 changes the value of $f$ from 0 to 1. Hence by fixing all variables in $x$ except for the $s(f)$ sensitive variables, we obtain the OR function on $s(f)$ variables. Since OR on $s(f)$ variables has $Q_0(\mbox{\rm OR})=s(f)$~\cite[Proposition~6.1]{bbcmw:polynomials}, it follows that $s(f)\leq Q_0(f)$. It is known (see for instance~\cite{nisan:pram&dt,bbcmw:polynomials}) that $D(f)\leq s(f)^2$ for monotone $f$, hence $D(f)\leq Q_0(f)^2$. \end{proof} Important examples of monotone functions are {\em AND-OR trees}. These can be represented as trees of depth $d$ where the $N$ leaves are the variables, and the $d$ levels of internal nodes are alternatingly labeled with ANDs and ORs. Using techniques from~\cite{bbcmw:polynomials}, it is easy to show that $Q_E(f)\geq N/2$ and $D(f)=N$ for such trees. However, we show that in the zero-error setting quantum computers can achieve significant speed-ups for such functions. These are in fact the first total functions with superlinear gap between quantum and classical zero-error complexity. Interestingly, the quantum algorithms for these functions are not just zero-error: if they output an answer $b\in\01$ then they also output a {\em $b$-certificate} for this answer. This is a set of indices of variables whose values force the function to the value $b$. We prove that for sufficiently large $d$, quantum computers can obtain near-quadratic speed-ups on $d$-level AND-OR trees which are uniform, i.e.~have branching factor $N^{1/d}$ at each level. Using the next lemma (which is proved in the appendix) we show that Theorem~\ref{thdvsqo} is almost tight: for every $\varepsilon>0$ there exists a total monotone $f$ with $Q_0(f)\in O(N^{1/2+\varepsilon})$. \begin{lemma}\label{lemandord} Let $d\geq 1$ and let $f$ denote the uniform $d$-level AND-OR tree on $N$ variables that has an OR as root. There exists a quantum algorithm $A_1$ that finds a 1-certificate in expected number of queries $O(N^{1/2+1/2d})$ if $f(x)=1$ and does not terminate if $f(x)=0$. Similarly, there exists a quantum algorithm $A_0$ that finds a 0-certificate in expected number of queries $O(N^{1/2+1/d})$ if $f(x)=0$ and does not terminate if $f(x)=1$. \end{lemma} \begin{theorem}\label{thmonsep} Let $d\geq 1$ and let $f$ denote the uniform $d$-level AND-OR tree on $N$ variables that has an OR as root. Then $Q_0(f)\in O(N^{1/2+1/d})$ and $R_2(f)\in\Omega(N)$. \end{theorem} \begin{proof} Run the algorithms $A_1$ and $A_0$ of Lemma~\ref{lemandord} side-by-side until one of them terminates with a certificate. This gives a certificate-finding quantum algorithm for $f$ with expected number of queries $O(N^{1/2+1/d})$. Run this algorithm for twice its expected number of queries and answer `don't know' if it hasn't terminated after that time. By Markov's inequality, the probability of non-termination is $\leq 1/2$, so we obtain an algorithm for our zero-error setting with $Q_0(f)\in O(N^{1/2+1/d})$ queries. The classical lower bound follows from combining two known results. First, an AND-OR tree of depth $d$ on $N$ variables has $R_0(f)\geq N/2^d$ \cite[Theorem~2.1]{hnw:readonce} (see also~\cite{saks&wigderson:trees}). Second, for such trees we have $R_2(f)\in\Omega(R_0(f))$~\cite{santha:montecarlo}. Hence $R_2(f)\in\Omega(N)$. \end{proof} This analysis is not quite optimal. It gives only trivial bounds for $d=2$, but a more refined analysis shows that we can also get speed-ups for such 2-level trees: \begin{theorem}\label{thgapqo} Let $f$ be the AND of $N^{1/3}$ ORs of $N^{2/3}$ variables each. Then $Q_0(f)\in \Theta(N^{2/3})$ and $R_2(f)\in\Omega(N)$. \end{theorem} \begin{proof} A similar analysis as before shows $Q_0(f)\in O(N^{2/3})$ and $R_2(f)\in\Omega(N)$. For the quantum lower bound: note that if we set all variables to 1 except for the $N^{2/3}$ variables in the first subtree, then $f$ becomes the OR of $N^{2/3}$ variables. This is known to have zero-error complexity exactly $N^{2/3}$~\cite[Proposition~6.1]{bbcmw:polynomials}, hence $Q_0(f)\in\Omega(N^{2/3})$. \end{proof} If we consider a tree with $\sqrt{N}$ subtrees of $\sqrt{N}$ variables each, we would get $Q_0(f)\in O(N^{3/4})$ and $R_2(f)\in\Omega(N)$. The best lower bound we can prove here is $Q_0(f)\in\Omega(\sqrt{N})$. However, if we also require the quantum algorithm to output a {\em certificate} for $f$, we can prove a tight quantum lower bound of $\Omega(N^{3/4})$. We do not give the proof here, which is a technical and more elaborate version of the proof of the classical lower bound of Theorem~\ref{thcertgap}. \section{Zero-error communication complexity} The results of the previous section can be translated to the setting of communication complexity~\cite{kushilevitz&nisan:cc}. Here there are two parties, Alice and Bob, who want to compute some relation $R\subseteq\01^N\times\01^N\times\01^M$. Alice gets input $x\in\01^N$ and Bob gets input $y\in\01^N$. Together they want to compute some $z\in\01^M$ such that $(x,y,z)\in R$, exchanging as few bits of communication as possible. The often studied setting where Alice and Bob want to compute some function $f:\01^N\times\01^N\rightarrow\01$ is a special case of this. In the case of {\em quantum} communication, Alice and Bob can exchange and process qubits, potentially giving them more power than classical communication. Let $g:\01^N\rightarrow\01$ be one of the AND-OR-trees of the previous section. We can derive from this a communication problem $f:\01^N\times\01^N\rightarrow\01$ by defining $f(x,y)=g(x\wedge y)$, where $x\wedge y\in\01^N$ is the vector obtained by bitwise AND-ing Alice's $x$ and Bob's $y$. Let us call such a problem a ``distributed'' AND-OR-tree. Buhrman, Cleve, and Wigderson~\cite{BuhrmanCleveWigderson98} show how to turn a $T$-query quantum black-box algorithm for $g$ into a communication protocol for $f$ with $O(T\log N)$ qubits of communication. Thus, using the upper bounds of the previous section, for every $\varepsilon>0$, there exists a distributed AND-OR-tree $f$ that has a $O(N^{1/2+\varepsilon})$-qubit zero-error protocol. It is conceivable that the classical zero-error communication complexity of these functions is $\omega(N^{1/2+\varepsilon})$; however, we are not able to prove such a lower bound at this time. Nevertheless, we are able to establish a quantum-classical separation for a relation that is closely related to the AND-OR-tree functions, which is explained below. For any AND-OR tree function $g : \01^N \rightarrow \01$ and input $x \in \01^N$, a {\em certificate for the value of $g$ on input $x$} is a subset $c$ of the indices $\{0,1,\ldots,N-1\}$ such that the values $\{x_i : i \in c\}$ determine the value of $g(x)$. It is natural to denote $c$ as an element of $\01^N$, representing the characteristic function of the set. For example, for \begin{equation} \label{cert_example} g(x_0,x_1,x_2,x_3) = (x_0 \vee x_1) \wedge (x_2 \vee x_3), \end{equation} a certificate for the value of $g$ on input $x = 1011$ is $c = 1001$, which indicates that $x_0 = 1$ and $x_3 = 1$ determine the value of $g$. We can define a communication problem based on finding these certificates as follows. For any AND-OR tree function $g : \01^N \rightarrow \01$ and $x, y \in \01^N$, a certificate for the value of $g$ on {\em distributed\/} inputs $x$ and $y$ is a subset $c$ of $\{0,1,\ldots,N-1\}$ (denoted as an element of $\01^N$) such that the values $\{(x_i,y_i) : i \in c\}$ determine the value of $g(x \wedge y)$. Define the relation $R\subseteq\01^N\times\01^N\times\01^N$ such that $(x,y,c)\in R$ iff $c$ is a certificate for the value of $g$ on distributed inputs $x$ and $y$. For example, when $R$ is with respect to the function $g$ of equation~(\ref{cert_example}), $(1011,1111,1001) \in R$, because, for $x=1011$ and $y=1111$, an appropriate certificate is $c=1001$. The zero-error certificate-finding algorithm for $g$ of the previous section, together with the \cite{BuhrmanCleveWigderson98}-translation from black-box algorithms to communication protocols, implies a zero-error quantum communication protocol for $R$. Thus, Theorem~\ref{thmonsep} implies that for every $\varepsilon>0$ there exists a relation $R\subseteq\01^N\times\01^N\times\01^N$ for which there is a zero-error quantum protocol with $O(N^{1/2+\varepsilon})$ qubits of communication. Although we suspect that the {\em classical} zero-error communication complexity of these relations is $\Omega(N)$, we are only able to prove lower bounds for relations derived from 2-level trees: \begin{theorem}\label{thcertgap} Let $g:\01^N\rightarrow\01$ be an AND of $N^{1/3}$ ORs of $N^{2/3}$ variables each. Let $R\subseteq\01^N\times\01^N\times\01^N$ be the certificate-relation derived from $g$. Then there exists a zero-error $O(N^{2/3}\log N)$-qubit quantum protocol for $R$, whereas, any zero-error classical protocol for $R$ needs $\Omega(N)$ bits of communication. \end{theorem} \begin{proof} The quantum upper bound follows from Theorem~\ref{thgapqo} and the \cite{BuhrmanCleveWigderson98}-reduction. For the classical lower bound, suppose we have a classical zero-error protocol $P$ for $R$ with $T$ bits of communication. We will show how we can use this to solve the Disjointness problem on $k=N^{1/3}(N^{2/3}-1)$ variables. (Given Alice's input $x\in\01^k$ and Bob's $y\in\01^k$, the Disjointness problem is to determine if $x$ and $y$ have a 1 at the same position somewhere.) Let $Q$ be the following classical protocol. Alice and Bob view their $k$-bit input as made up of $N^{1/3}$ subtrees of $N^{2/3}-1$ variables each. They add a dummy variable with value 1 to each subtree and apply a random permutation to each subtree (Alice and Bob have to apply the {\em same} permutation to a subtree, so we assume a public coin). Call the $N$-bit strings they now have $x'$ and $y'$. Then they apply $P$ to $x'$ and $y'$. Since $f(x',y')=1$, after an expected number of $O(T)$ bits of communication $P$ will deliver a certificate which is a common 1 in each subtree. If one of these common 1s is non-dummy then Alice and Bob output 1, otherwise they output 0. It is easy to see that this protocol solves Disjointness with success probability 1 if $x\wedge y=\vec{0}$ and with success probability $\geq 1/2$ if $x\wedge y\neq\vec{0}$. It assumes a public coin and uses $O(T)$ bits of communication. Now the well-known $\Omega(k)$ bound for classical bounded-error Disjointness on $k$ variables~\cite{ks:disj,razborov:disj} implies $T\in\Omega(k)=\Omega(N)$. \end{proof} The relation of Theorem~\ref{thcertgap} is ``total'', in the sense that, for every $x, y \in \01^N$, there exists a $c$ such that $(x,y,c) \in R$. It should be noted that one can trivially construct a total relation from any partial function by allowing any output for inputs that are outside the domain of the function. In this manner, a total relation with an exponential quantum-classical zero-error gap can be immediately obtained from the distributed Deutsch-Jozsa problem of~\cite{BuhrmanCleveWigderson98}. The total relation of Theorem~\ref{thcertgap} is different from this in that it is not a trivial extension of a partial function. After reading a first version of this paper, Hartmut Klauck proved a separation which is the first example of a total {\em function\/} with superlinear gap between quantum and classical zero-error communication complexity~\cite{klauck:qpcom}. Consider the iterated non-disjointness function: Alice and Bob each receive $s$ sets of size $n$ from a size-$poly(n)$ universe (so the input length is $N\in\Theta(sn\log n)$ bits), and they have to output 1 iff all $s$ pairs of sets intersect. Klauck's function $f$ is an intricate subset of this iterated non-disjointness function, but still an explicit and total function. Results of~\cite{hromkovic&schnitger:limadvice} about limited non-deterministic communication complexity imply a lower bound for classical zero-error protocols for $f$. On the other hand, because $f$ can be written as a 2-level AND-OR-tree, the methods of this paper imply a more efficient quantum zero-error protocol. Choosing $s=n^{5/6}$, Klauck obtains a polynomial gap: \begin{theorem}[Klauck \cite{klauck:qpcom}]\label{thklauck} For $N\in\Theta(n^{11/6}\log n)$ there exists a total function $f:\01^N\times\01^N\rightarrow\01$, such that there is a quantum zero-error protocol for $f$ with $O(N^{10/11+\varepsilon})$ qubits of communication (for all $\varepsilon>0$), whereas every classical zero-error protocol for $f$ needs $\Omega(N/\log N)$ bits of communication. \end{theorem} \section{Quantum complexity of graph properties}\label{secgraphprop} {\em Graph properties} form an interesting subset of the set of all Boolean functions. Here an input of $N=n(n-1)$ bits represents the edges of a directed graph on $n$ vertices. (Our results hold for properties of {\em directed} as well as {\em undirected} graphs.) A {\em graph property} $P$ is a subset of the set of all graphs that is closed under permutation of the nodes (so if $X,Y$ represent isomorphic graphs, then $X\in P$ iff $Y\in P$). We are interested in the number of queries of the form ``is there an edge from node $i$ to node $j$?'' that we need to determine for a given graph whether it has a certain property $P$. Since we can view $P$ as a total function on $N$ variables, we can use the notations $D(P)$, etc. A property $P$ is {\em evasive} if $D(P)=n(n-1)$, so if in the worst case all $N$ edges have to be examined. The complexity of graph properties has been well-studied classically, especially for {\em monotone} graph properties (a property is monotone if adding edges cannot destroy the property). In the sequel, let $P$ stand for a (non-constant) monotone graph property. Much research revolved around the so-called Aanderaa-Karp-Rosenberg conjecture or {\em evasiveness conjecture}, which states that every $P$ is evasive. This conjecture is still open; see~\cite{lovasz:graphprop} for an overview. It has been proved for $n$ equals a prime power~\cite{kss:topo}, but for other $n$ the best known general bound is $D(P)\in\Omega(n^2)$ \cite{rivest&vuillemin,kss:topo,king:graphprop}. (Evasiveness has also been proved for {\em bipartite} graphs~\cite{yao:bipartite}.) For the classical zero-error complexity, the best known general result is $R_0(P)\in\Omega(n^{4/3})$~\cite{hajnal:graphprop}, but it has been conjectured that $R_0(P)\in\Theta(n^2)$. To the best of our knowledge, no $P$ is known to have $R_2(P)\in o(n^2)$. In this section we examine the complexity of monotone graph properties on a quantum computer. First we show that if we replace exact classical algorithms by exact quantum algorithms, then the evasiveness conjecture fails. However, the conjecture does hold up to a constant factor. \begin{theorem} For all $P$, $Q_E(P)\in\Omega(n^2)$. There is a $P$ such that $Q_E(P)<n(n-1)$ for every $n>2$. \end{theorem} \begin{proof} For the lower bound, let $deg(f)$ denote the degree of the unique multilinear multivariate polynomial $p$ that represents a function $f$ (i.e.~$p(X)=f(X)$ for all $X$). \cite{bbcmw:polynomials} proves that $Q_E(f)\geq deg(f)/2$ for every $f$. Dodis and Khanna~\cite[Theorem~5.1]{dodis&khanna} prove that $deg(P)\in\Omega(n^2)$ for all monotone graph properties $P$. Combining these two facts gives the lower bound. Let $P$ be the property ``the graph contains more than $n(n-1)/2$ edges''. This is just a special case of the Majority function. Let $f$ be Majority on $N$ variables. It is known that $Q_E(f)\leq N+1-e(N)$, where $e(N)$ is the number of 1s in the binary expansion of $N$. This was first noted by Hayes, Kutin and Van Melkebeek~\cite{hkm:qmajority}. It also follows immediately from classical results~\cite{saks&werman:majority,ars:majority} that show that an item with the Majority value can be identified classically deterministically with $N-e(N)$ {\em comparisons} between bits (a comparison between two black-box-bits is the XOR of two bits, which can be computed with 1 quantum query~\cite{cemm:revisited}). One further query to this item suffices to determine the Majority value. For $N=n(n-1)$ and $n>2$ we have $e(N)\geq 2$ and hence $Q_E(f)\leq N-e(N)+1<N$. \end{proof} In the zero-error case, we can show polynomial gaps between quantum and classical complexities, so here the evasiveness conjecture fails even if we ignore constant factors. \begin{theorem} For all $P$, $Q_0(P)\in\Omega(n)$. There is a $P$ such that $Q_0(P)\in O(n^{3/2})$ and $R_2(P)\in\Omega(n^2)$. \end{theorem} \begin{proof} The quantum lower bound follows from $D(P)\leq Q_0(P)^2$ (Theorem~\ref{thdvsqo}) and $D(P)\in\Omega(n^2)$. Consider the property ``the graph contains a star'', where a star is a node that has edges to all other nodes. This property corresponds to a 2-level tree, where the first level is an OR of $n$ subtrees, and each subtree is an AND of $n-1$ variables. The $n-1$ variables in the $i$th subtree correspond to the $n-1$ edges $(i,j)$ for $j\neq i$. The $i$th subtree is 1 iff the $i$th node is the center of a star, so the root of the tree is 1 iff the graph contains a star. Now we can show $Q_0(P)\in O(n^{3/2})$ and $R_2(P)\in\Omega(n^2)$ analogously to Theorem~\ref{thgapqo}. \end{proof} Combined with the translation of a quantum algorithm to a polynomial \cite{bbcmw:polynomials}, this theorem shows that a ``zero-error polynomial'' for the STAR-graph property can have degree $O(n^{3/2})$. Thus proving a general lower bound on zero-error polynomials for graph properties will not improve Hajnal's randomized lower bound of $n^{4/3}$ further then $n^{3/2}$. In particular, a proof that $R_0(P)\in\Omega(n^2)$ cannot be obtained via a lower bound on degrees of polynomials. This contrasts with the case of exact computation, where the $\Omega(n^2)$ lower bound on $deg(P)$ implies both $D(P)\in\Omega(n^2)$ and $Q_E(P)\in\Omega(n^2)$. Finally, for the bounded-error case we have quadratic gaps between quantum and classical: the property ``the graph has at least one edge'' has $Q_2(P)\in O(n)$ by Grover's quantum search algorithm. Combining that $D(P)\in\Omega(n^2)$ for all $P$ and $D(f)\in O(Q_2(f)^4)$ for all monotone $f$~\cite{bbcmw:polynomials}, we also obtain a general lower bound: \begin{theorem} For all $P$, we have $Q_2(P)\in\Omega(\sqrt{n})$. There is a $P$ such that $Q_2(P)\in O(n)$. \end{theorem} \noindent {\large\bf Acknowledgments}\\ We thank Hartmut Klauck for informing us about Theorem~\ref{thklauck}, Ramamohan Paturi for Lemma~\ref{lemchebbound}, David Deutsch, Wim van Dam, and Michele Mosca for helpful discussions which emphasized the importance of small-error quantum search, and Mosca and Yevgeniy Dodis for helpful discussions about graph properties.
\section{ Introduction } The quantized motional states of atoms or ions in confining potentials offer interesting possibilities for a variety of applications, such as the preparation and study of nonclassical (i.e., manifestly quantum) states \cite{Meekhof96,Monroe96,Leibfried96,Cirac96,Deutsch98}, and the storage and manipulation of quantum information (e.g., ``qubits''), with particular reference to quantum logic operations and quantum computing \cite{Cirac95a,Monroe95,King98,Turchette98a,Hughes96,Steane97,Wineland98}. These possibilities stem from the relatively long coherence times that can be achieved with motional states (due to the absence of strong damping mechanisms) and the precision with which transformations between motional states can be controlled using laser-light-induced transitions. However, while motional states are well-suited to the storage and manipulation of quantum states, for the communication of quantum information from one physical location to another it is clear that photons are the preferred carriers of the information. For this reason, it is necessary to provide and examine configurations in which motional states can be efficiently and reliably transferred to states of light, and vice-versa. Here enters the field of cavity quantum electrodynamics (cavity QED); in particular, configurations in which a single mode of the electromagnetic field supported by an optical cavity is strongly coupled to a transition in a single atom. It is possible, via the internal atomic transition, to also couple the cavity field to the external (quantized) motion of the trapped atom or ion \cite{Zeng94,Buzek97,Harrison97}, and in this work we will examine such a coupling that enables the above-mentioned state transfer. \section{ Model } Our model consists of a single two-level atom (or ion) confined in a harmonic trap located inside an optical cavity. The atomic transition of frequency $\omega_{\rm a}$ is coupled to a single mode of the cavity field of frequency $\omega_{\rm c}$ and is also assumed to be driven by an external (classical) laser field of frequency $\omega_{\rm L}$ -- the cavity and laser field frequencies will be chosen so as to drive Raman transitions that couple neighbouring vibrational levels of the external motion. The physical setup and excitation scheme are depicted in Fig.~1. The cavity is aligned along the $x$-axis, while the laser field is incident from a direction in the $y$-$z$ plane (i.e., perpendicular to the $x$-axis). The Hamiltonian describing the internal and external atomic degrees of freedom plus the atom-cavity and atom-laser couplings takes the form (in a frame rotating at the laser frequency) \begin{eqnarray} \hat{H}_0 = && \sum_{j=x,y,z} \hbar\nu_j (\hat{b}_j^\dagger\hat{b}_j+1/2) + \hbar\delta \hat{a}^\dagger\hat{a} + \hbar\Delta\hat{\sigma}_+\hat{\sigma}_- \nonumber \\ && \;\;\; +\, \hbar \left[ {\cal E}_{\rm L}(\hat{y},\hat{z},t) \hat{\sigma}_+ + {\cal E}_{\rm L}(\hat{y},\hat{z},t)^\ast \hat{\sigma}_- \right] \nonumber \\ && \;\;\; +\, \hbar g_0 \sin (k\hat{x}) (\hat{a}^\dagger \hat{\sigma}_- + \hat{\sigma}_+\hat{a} ) . \end{eqnarray} Here, $\{\nu_x,\nu_y,\nu_z\}$ are the harmonic oscillation frequencies along the principal axes of the trap, $\hat{b}_j$ and $\hat{a}$ are annihilation operators for the quantized atomic motion and cavity field, respectively, $\hat{\sigma}_- =|g\rangle\langle e|$ is the atomic lowering operator, and $\delta =\omega_{\rm c}-\omega_{\rm L}$ and $\Delta =\omega_{\rm a}-\omega_{\rm L}$. The quantity ${\cal E}_{\rm L}(\hat{y},\hat{z},t)$ is the (possibly time-dependent) amplitude of the laser field; note again that we assume that this field has no spatial dependence along the $x$ direction. Finally, the single-photon atom-cavity dipole coupling strength is given by $g_0$, while the sine function describes the standing wave structure of the cavity field (we assume that the centre of the trap is located at a {\em node} of the cavity field), with $k=2\pi /\lambda$ the wavenumber of the field and $\hat{x}=[\hbar /(2m\nu_x)]^{1/2} (\hat{b}_x+\hat{b}_x^\dagger )$. Allowing for cavity damping and atomic spontaneous emission, quantum Langevin equations for the system operators can be derived straightforwardly. For the moment, however, we will ignore the effects of atomic spontaneous emission on the grounds that the detunings of the laser and cavity fields from the atomic transition frequency are very large, and hence that population of the excited atomic state $|e\rangle$ is negligible (we will return to the effects of spontaneous emission in the discussion at the end of the paper). On this basis, we are also able to adiabatically eliminate the internal atomic dynamics from the problem. We will also ignore any forms of motional decoherence or heating associated with imperfections in the trap itself \cite{Wineland98} on the basis that such effects occur on a timescale slow compared with the operations we will be considering. Again, we will return to this point and discuss it more quantitatively at the end of the paper. Finally, we assume that the size of the harmonic trap (in all directions) is small compared to the optical wavelength; under these conditions, we can make the approximation $\sin (k\hat{x})\simeq \eta_x (\hat{b}_x+\hat{b}_x^\dagger )$, where $\eta_x$ ($\ll 1$) is the Lamb-Dicke parameter. Given this assumption, it is also possible to design a configuration for which we can neglect all position dependence in the laser field \cite{laserantinode}; that is, we can assume a situation where ${\cal E}_{\rm L}(\hat{y},\hat{z},t)\simeq {\cal E}_{\rm L}(t)e^{-i\phi_{\rm L}}$ [with ${\cal E}_{\rm L}(t)$ a real quantity]. Henceforth, the problem essentially becomes one-dimensional and we can restrict our attention to just the $x$ direction. To first order in $\eta_x$, equations of motion for the operators $\hat{a}$ and $\hat{b}_x$ then follow as \begin{eqnarray} \dot{\hat{a}} &=& -(\kappa +i\delta )\hat{a} + i \frac{g_0\eta_x {\cal E}_{\rm L} (t)e^{-i\phi_{\rm L}}}{\Delta} \, (\hat{b}_x+\hat{b}_x^\dagger ) \nonumber \\ && \;\;\;\;\;\; - \sqrt{2\kappa}\, \hat{a}_{\rm in}(t) \; , \\ \dot{\hat{b}}_x &=& -i\nu_x\hat{b}_x + i\frac{g_0\eta_x {\cal E}_{\rm L}(t)}{\Delta} \, \left( \hat{a}^\dagger e^{-i\phi_{\rm L}} + \hat{a} e^{i\phi_{\rm L}} \right) \; , \end{eqnarray} where $\kappa$ is the decay rate of the cavity field and $\hat{a}_{\rm in}(t)$ is a quantum noise operator describing the input to the cavity field and satisfying the commutation relation $[\hat{a}_{\rm in}(t),\hat{a}_{\rm in} ^\dagger (t^\prime )] =\delta (t-t^\prime )$. We now make the transformation $\hat{b}_x=e^{-i\nu_x t}\tilde{b}_x$, $\hat{a}=e^{-i\nu_x t}\tilde{a}$, and choose $\delta =\nu_x$ (i.e., tune to the ``first lower sideband''). Assuming that $\nu_x\gg\kappa$, $|(g_0\eta_x /\Delta ){\cal E}_{\rm L}(t)|$, and $|\dot{{\cal E}}_{\rm L}(t)/{\cal E}_{\rm L}(t)|$, the oscillating terms in the resulting equations may be dropped in a rotating-wave approximation to yield \begin{eqnarray} \dot{\tilde{a}} &=& -\kappa \tilde{a} + i\frac{g_0\eta_x {\cal E}_{\rm L}(t) e^{-i\phi_{\rm L}}}{\Delta} \, \tilde{b}_x - \sqrt{2\kappa}\, \tilde{a}_{\rm in}(t) \\ \dot{\tilde{b}}_x &=& i\frac{g_0\eta_x {\cal E}_{\rm L}(t)}{\Delta} \tilde{a} e^{i\phi_{\rm L}} . \end{eqnarray} These equations simply describe a pair of coupled harmonic oscillators, one of which is damped. In terms of a Hamiltonian, the coupling is given by [defining $\Omega (t)=-g_0\eta_x {\cal E}_{\rm L}(t)/\Delta$ (real)] \begin{equation} \tilde{H}(t) = \hbar \Omega (t) \left( \tilde{a}^\dagger \tilde{b}_x e^{-i\phi_{\rm L}} + \tilde{b}_x^\dagger \tilde{a} e^{i\phi_{\rm L}} \right) \; , \end{equation} a result derived by Zeng and Lin \cite{Zeng94}. \section{ Quantum state transfer } As pointed out by Zeng and Lin \cite{Zeng94}, when cavity losses can be neglected the above coupling enables complete (pure or mixed) state transfer between the atomic motion and the cavity light field. For example, given a finite laser pulse duration, and assuming for simplicity that ${\cal E}_{\rm L}(t)$ is a real and positive function of time, then solutions for $\tilde{a}(t=+\infty )\equiv\tilde{a}(+\infty )$ and $\tilde{b}_x(t=+\infty )\equiv\tilde{b}_x(+\infty )$ can be derived as \begin{eqnarray} \tilde{a}(+\infty ) &=& \tilde{a}(-\infty ) \cos\theta - ie^{-i\phi_{\rm L}}\tilde{b}_x(-\infty ) \sin\theta \\ \tilde{b}_x(+\infty ) &=& - ie^{i\phi_{\rm L}}\tilde{a}(-\infty ) \sin\theta + \tilde{b}_x(-\infty ) \cos\theta \; , \end{eqnarray} where $\theta =\int_{-\infty}^\infty \Omega (\tau )d\tau$. Choosing $\theta =(N+1/2)\pi$, with $N$ an integer, and $\phi_{\rm L} =\pi /2$ yields \begin{equation} \tilde{a}(+\infty ) = \mp \tilde{b}_x(-\infty ) \; , \;\;\;\; \tilde{b}_x(+\infty ) = \pm \tilde{a}(-\infty ) \; , \end{equation} from which it follows that, given an initial vacuum state of the field, {\em any initial state of the motion can be transferred one-to-one to the state of the cavity field} and the motion is reduced to its ground state (for other examples of this kind of state exchange between harmonic oscillator modes, see \cite{Steinbach97,Heinzen90}). \subsection{ Underdamped regime } The result derived above demonstrating the possibility of complete state transfer between the quantized atomic motion and the cavity field mode obviously offers some very interesting further possibilities. If the damping of the cavity mode is sufficiently weak, then one can imagine a situation in which a suitable laser pulse is applied so as to transfer a motional state to the cavity mode (in a time short compared to $\kappa^{-1}$), after which the cavity field is allowed to decay. Making homodyne measurements on the output field from the cavity and using the well-established method of optical homodyne tomography \cite{Vogel89,Smithey93}, the density matrix of the cavity field mode, and hence of the initial motional state, could be reconstructed from many repeated cycles of preparation and measurement. Given that the motion is always left in its ground state after the laser pulse, this can also be seen as a novel means of cooling, or ``resetting,'' the atomic motion in a single operation \cite{Heinzen90b}. Of course, in the regime of operation that we are assuming (the resolved sideband limit), conventional sideband cooling (using atomic spontaneous emission) would also be an efficient means of cooling the motion to the ground state. \subsection{ Overdamped regime } The opposite limit, in which $\kappa$ is large compared to the magnitude of the effective coupling rate $\Omega (t)$ (but, of course, still small compared to the trap frequency $\nu_x$) is actually of more interest to us and indeed allows further simplification of the model. In particular, we can consider adiabatically eliminating the cavity mode from the dynamics, i.e., setting $\dot{\tilde{a}}=0$ and substituting \begin{equation} \label{adiab_elim} \tilde{a} \simeq -\,\frac{\Omega (t)}{\kappa} \, \tilde{b}_x - \sqrt{2/\kappa}\, \tilde{a}_{\rm in}(t) \end{equation} into the equation for $\dot{\tilde{b}}_x$ to give \begin{eqnarray} \dot{\tilde{b}}_x &\simeq & - \frac{\left[ \Omega (t) \right]^2}{\kappa} \tilde{b}_x + \Omega (t) \sqrt{2/\kappa}\, \tilde{a}_{\rm in}(t) \nonumber \\ & \equiv & - \Gamma (t) \tilde{b}_x + \sqrt{2\Gamma (t)} \, \tilde{a}_{\rm in}(t)\; , \label{eq:btilde} \end{eqnarray} where we have set $\phi_{\rm L}=\pi /2$ for simplicity. This equation simply describes a quantum harmonic oscillator subject to damping at the (possibly time-dependent) rate $\Gamma (t)$. In the case of a vacuum cavity input field, it obviously models sideband cooling to the ground state due to a form of cavity-induced spontaneous emission \cite{Cirac95b,Blatt99}. \subsubsection{ Driven cavity: light-to-motion state transfer } However, one can also consider different kinds of inputs to the cavity field, i.e., choices of the input field operator $\tilde{a}_{\rm in}(t)$ that give rise to nontrivial input field statistics. This is of interest because, given the simple linear form of (\ref{eq:btilde}), it follows that the statistics of the input field can be ``written onto'' the state of the oscillator. In particular, assuming that $\Gamma (t)=\Gamma$, a constant, then in frequency space the solution to (\ref{eq:btilde}) is simply \begin{equation} \tilde{b}_x(\omega ) = \frac{\sqrt{2\Gamma}\, \tilde{a}_{\rm in}(\omega )}{i\omega -\Gamma} \, . \end{equation} For example, the input field could be an {\em ideal quantum squeezed vacuum} as derived from the output of a degenerate parametric amplifier \cite{Kimble92,Turchette98b}. If the squeezing happens to be broadband (with respect to the characteristic rates associated with the system upon which it is incident), then the appropriate input field correlation functions can be written in the forms \cite{Gardiner91} \begin{eqnarray} \langle \tilde{a}_{\rm in}^\dagger (\omega )\tilde{a}_{\rm in}(\omega^\prime ) \rangle &=& N \, \delta (\omega -\omega^\prime ) \\ \langle \tilde{a}_{\rm in}(\omega ) \tilde{a}_{\rm in}(\omega^\prime ) \rangle &=& M \, \delta (\omega +\omega^\prime ) \, , \end{eqnarray} with $M=|M|e^{i\theta}$ and $|M|^2=N(N+1)$. Given such an input, the system is equivalently described by the master equation \cite{Gardiner91} \begin{eqnarray} \label{eq:drhomdt} \dot{\rho}_{\rm m} = && \Gamma (N+1)(2\tilde{b}_x\rho_{\rm m}\tilde{b}_x^\dagger - \tilde{b}_x^\dagger\tilde{b}_x\rho_{\rm m} - \rho_{\rm m}\tilde{b}_x^\dagger\tilde{b}_x) \nonumber \\ && + \Gamma N(2\tilde{b}_x^\dagger\rho_{\rm m}\tilde{b}_x - \tilde{b}_x\tilde{b}_x^\dagger\rho_{\rm m} - \rho_{\rm m}\tilde{b}_x\tilde{b}_x^\dagger ) \nonumber \\ && - \Gamma M(2\tilde{b}_x\rho_{\rm m}\tilde{b}_x - \tilde{b}_x\tilde{b}_x\rho_{\rm m} - \rho_{\rm m}\tilde{b}_x\tilde{b}_x) \nonumber \\ && - \Gamma M^\ast (2\tilde{b}_x^\dagger\rho_{\rm m}\tilde{b}_x^\dagger - \tilde{b}_x^\dagger\tilde{b}_x^\dagger\rho_{\rm m} - \rho_{\rm m}\tilde{b}_x^\dagger \tilde{b}_x^\dagger ) \, , \end{eqnarray} where $\rho_{\rm m}$ is the density operator for the motion of the trapped atom. In steady state, the density operator is that of an ideal squeezed state, that is \begin{equation} \rho_{\rm m}^{\rm ss} = \hat{S}|0\rangle\langle 0|\hat{S}^+ \, , \end{equation} where $\hat{S}$ is the squeezing operator \cite{Walls94}, i.e., $\hat{S}\tilde{b}\hat{S}^+=\mu\tilde{b}_x +\nu\tilde{b}_x^\dagger$ with $\mu =(N+1)^{1/2}$ and $\nu =N^{1/2}e^{i\theta}$. There are of course other ways of preparing such nonclassical states of the motion which have indeed already been implemented experimentally \cite{Meekhof96,Monroe96,Leibfried96}. These preparations have typically employed pulsed {\em classical} light fields to facilitate the required motional state transformations. The above scheme is novel in that it involves the direct transfer of statistics from a {\em nonclassical} continuous-wave light field to the motional state of the trapped atom. \subsubsection{ Numerical calculations } To numerically model state transfer between light and motion in the overdamped regime described above, we consider, as above, the cavity mode to be resonantly driven by squeezed light from a degenerate parametric oscillator, as illustrated in Fig.~2. For our simulations, we include the dynamics of the parametric oscillator using the cascaded systems formalism developed in \cite{Gardiner93,Carmichael93} (further applications of the formalism are given in \cite{Gardiner94}). In particular, we model our system with the master equation \begin{eqnarray} \label{eq:sqme} \dot{\rho} &=& - \frac{i}{\hbar} [\tilde{H}_{ab}(t)+\tilde{H}_c, \rho ] \nonumber \\ && +\, \kappa_a \left( 2\tilde{a}\rho\tilde{a}^\dagger - \tilde{a}^\dagger\tilde{a}\rho - \rho\tilde{a}^\dagger\tilde{a} \right) \nonumber \\ && +\, \kappa_c \left( 2\tilde{c}\rho\tilde{c}^\dagger - \tilde{c}^\dagger\tilde{c}\rho - \rho\tilde{c}^\dagger\tilde{c} \right) \nonumber \\ && -\, 2\sqrt{\kappa_a\kappa_c} \, \left( [\tilde{a}^\dagger , \tilde{c}\rho ] + [\rho\tilde{c}^\dagger ,\tilde{a}] \right) \, . \end{eqnarray} Here, $\tilde{H}_{ab}(t)$ describes the effective coupling between the vibrational motion of the trapped atom and the cavity light field as derived above. However, in addition we retain the rotating, or nonsecular terms, i.e., we take \begin{eqnarray} \tilde{H}_{ab}(t) &=& \hbar \Omega \left( \tilde{a}^\dagger \tilde{b}_x e^{-i\phi_{\rm L}} + \tilde{b}_x^\dagger \tilde{a} e^{i\phi_{\rm L}} \right) \nonumber \\ &+& \, \hbar \Omega \left( \tilde{a}^\dagger \tilde{b}_x^\dagger e^{-i\phi_{\rm L}+2i\nu_x t} + \tilde{a} \tilde{b}_x e^{i\phi_{\rm L}-2i\nu_x t} \right) \, . \end{eqnarray} The coupling parameter $\Omega$ is assumed to be a constant. The Hamiltonian $\tilde{H}_c$ models the parametric oscillator (driven below threshold), taking the form \begin{equation} \tilde{H}_c = \frac{1}{2} i\hbar \left[ \epsilon^\ast\tilde{c}^2 -\epsilon (\tilde{c}^\dagger )^2 \right] \, , \end{equation} where $\tilde{c}$ is the annihilation operator for the cavity mode of the parametric oscillator and $\epsilon =|\epsilon |e^{i\theta}$ is the amplitude of the coherent field driving the oscillator. The linewidth of the (one-sided) parametric oscillator cavity mode is $\kappa_c$. Finally, the last term in (\ref{eq:sqme}) describes the (unidirectional) coupling of the incoming squeezed light to the cavity mode. This coupling is assumed to be ideal. By using a truncated state basis, the master equation (\ref{eq:sqme}) is numerically propagated until a steady state is achieved. In fact, due to the time-dependent terms in $\tilde{H}_{ab}(t)$, only a quasi-steady state can in principle be achieved, but for the parameters we consider only a very weak time-dependence (i.e., a weak modulation) occurs. The elements of the reduced density matrix of the motional mode in the steady state ($\rho_{\rm m}^{\rm sim}$) are shown in Fig.~3 for the choice of parameters $\nu_x/\kappa_a=10$, $\Omega /\kappa_a =0.1$, $\kappa_c/\kappa_a=1$ and $\epsilon /\kappa_c=0.3$ (corresponding to 71\% maximum squeezing in the input light field). These parameters are in the regime for which the master equation (\ref{eq:drhomdt}) should be valid and, indeed, we find the steady state of the motion to be approached on a timescale $\Gamma^{-1}=(\Omega^2/\kappa_a)^{-1}=100\kappa_a^{-1}$. Characteristic squeezed state features are evident in the figure (e.g., only even number states are populated) and the fidelity with which the predicted ideal squeezed state is achieved is computed to be $\langle 0|\hat{S}^+\rho_{\rm m}^{\rm sim}\hat{S}|0\rangle \simeq 0.99$ for the appropriate values of $N$ and $M$ in the theory. \subsection{ Entanglement transfer from light fields to separated trapped atoms } The scheme outlined above can be extended to the transfer of quantum mechanical entanglement from light fields to motional states of two or more trapped atoms at physically separated sites. Consider, for example, the pair of output fields from a {\em nondegenerate parametric amplifier} (the fields may be nondegenerate in polarization or in frequency) \cite{Kimble92}. At the output from the parametric amplifier, these fields could be separated in space and then made to impinge upon two cavities containing trapped atoms in the configuration described above. The quantum mechanical correlations that exist between the two light fields generated by parametric downconversion could thus be transferred (in steady state) to correlations between motional states of trapped atoms at two distinct sites. \subsection{ Quantum teleportation of motional states } An exciting recent development in the field of quantum communication has been the experimental investigation of schemes for the teleportation of quantum states \cite{Boschi98,Bouwmeester97,Furusawa98}. Of particular interest in the present context is the demonstration by Furusawa {\em et al}. \cite{Furusawa98} of unconditional quantum teleportation of optical coherent states using squeezed-state fields and entanglement of the sort discussed above. By incorporating cavities containing trapped ions in the state transfer configuration of this work (in the bad cavity limit), it should be possible to employ the scheme of \cite{Furusawa98} for the teleportation of {\em motional} states. In particular, a motional state could be ``mapped'' onto a light field which enters the configuration of \cite{Furusawa98} at the ``sending'' station. This field is teleported to a ``receiving'' station where it is made to impinge upon a second trapped-ion plus cavity system in the state transfer configuration. The teleported light field is thus mapped onto the motional mode of the second trapped ion and teleportation of the motional state is completed. We will examine teleportation of motional states in more detail in a future work. \subsection{ Motion-to-light state transfer: generation of nonclassical output light fields } Given the variety of, and efficiency with which, nonclassical motional states of single trapped atoms have been experimentally realized \cite{Meekhof96,Monroe96,Leibfried96}, it is worth noting the potential of our scheme as a {\em source} of nonclassical output light fields. The output light field is related to the input and internal cavity fields by \cite{Gardiner91,Walls94} \begin{equation} \tilde{a}_{\rm out}(t) = \tilde{a}_{\rm in}(t) + \sqrt{2\kappa}\, \tilde{a}(t) \, , \end{equation} which in the overdamped limit becomes [using (\ref{adiab_elim})] \begin{equation} \tilde{a}_{\rm out}(t) \simeq - \tilde{a}_{\rm in}(t) - \sqrt{2\Gamma (t)}\, \tilde{b}_x(t) \, , \end{equation} where we assume that $\Omega (t)\geq 0$. Hence, given a vacuum field input to the cavity, the output field is determined by the motional state of the trapped atom. Further, depending on the nature of the motional state preparation, the output may be pulsed {\em or} continuous; for a continuous output one would have $\Gamma (t)=\Gamma$, a constant, and the motional state preparation scheme would have to operate in a continuous manner also. As an example, consider squeezed motional states, which may be generated by applying an electric field gradient with a frequency $2\nu_x$ to the ion \cite{Heinzen90}, or by irradiating the ion with two laser beams differing in frequency by $2\nu_x$ \cite{Meekhof96}. Note that in \cite{Meekhof96} squeezed states of the motion were produced exhibiting a reduction in the variance of the squeezed quadrature by a factor of 40. Such quadrature noise reduction has yet to be approached via traditional optical means, suggesting that the present state transfer configuration is worthy of further investigation \cite{LDlimit}. \section{ Transfer of a motional state between separated trapped atoms } Recently, Cirac {\em et al.} \cite{Cirac97} (see also \cite{vanEnk97,Pellizzari97,vanEnk98}) demonstrated how quantum transmission of a qubit between two nodes of a quantum network can be implemented in a physical system using light as the carrier of the quantum information. In particular, they showed how the transformation \begin{equation} \label{qubittrans} (c_0|0\rangle_1 + c_1|1\rangle_1) \otimes |0\rangle_2 \rightarrow |0\rangle_1 \otimes (c_0|0\rangle_2 + c_1|1\rangle_2) \end{equation} can be achieved where $|0\rangle_1$ and $|1\rangle_1$ are internal states of an atom at node 1 and $|0\rangle_2$ and $|1\rangle_2$ are the corresponding states of a second atom at (the spatially separated) node 2. At each node, the atom is located within a cavity supporting a single mode of the electromagnetic field, with which it is made to undergo a controlled time-dependent interaction via a laser-assisted Raman process. With suitably chosen laser pulses at the two nodes, the transmission described by (\ref{qubittrans}) can be faithfully reproduced, facilitated by the transfer of a {\em photon} wave packet between the two nodes. In the same spirit, we consider here the transmission of {\em arbitrary motional states} of trapped atoms between two distinct sites, facilitated once again by cavity light fields and photon wave packets. We consider two separated atom-cavity arrangements, each in the configuration described in Section II, so that the combined system Hamiltonian can be written as \begin{eqnarray} \tilde{H} &=& \hbar \Omega_1 (t) \left( \tilde{a}_1^\dagger \tilde{b}_1 e^{-i\phi_1} + \tilde{b}_1^\dagger \tilde{a}_1 e^{i\phi_1} \right) \nonumber \\ && + \hbar \Omega_2 (t) \left( \tilde{a}_2^\dagger \tilde{b}_2 e^{-i\phi_2} + \tilde{b}_2^\dagger \tilde{a}_2 e^{i\phi_2} \right) \, , \end{eqnarray} where the subscripts $\{ 1,2\}$ denote the site of each atom or cavity mode and we now omit the subscript $x$ for simplicity. \subsection{ Cascaded systems model } The two cavity modes are each damped at rate $\kappa$, but coupling between their external fields is assumed to be unidirectional. In particular, the output from cavity 1 is incident upon (i.e., provides the input field to) cavity 2, but not {\em vice versa}. Such a situation is depicted in Fig.~4 and is again modeled by the cascaded systems formalism introduced earlier \cite{Gardiner93,Carmichael93}. In this formalism, the master equation for our system is derived in the form \begin{equation} \dot{\rho} = \left( {\cal L}_0 + {\cal L}_{\rm c} \right) \rho \, , \end{equation} where ${\cal L}_0\rho = -(i/\hbar )[\tilde{H},\rho ]$ and \begin{eqnarray} \label{Lc} {\cal L}_{\rm c}\rho &=& \kappa \left( 2\tilde{a}_1\rho\tilde{a}_1^\dagger - \tilde{a}_1^\dagger\tilde{a}_1\rho - \rho\tilde{a}_1^\dagger \tilde{a}_1 \right) \nonumber \\ && + \kappa \left( 2\tilde{a}_2\rho\tilde{a}_2^\dagger - \tilde{a}_2^\dagger\tilde{a}_2\rho - \rho\tilde{a}_2^\dagger \tilde{a}_2 \right) \nonumber \\ && - 2\kappa \left( [\tilde{a}_2^\dagger ,\tilde{a}_1\rho ] + [\rho\tilde{a}_1^\dagger ,\tilde{a}_2 ] \right) , \end{eqnarray} where a vacuum input to the first cavity has been assumed. The last term in (\ref{Lc}) provides the desired unidirectional coupling in the theory. To simplify the model, we assume, as before, that the decay rate $\kappa$ is large compared to other rates in the system (apart from the trap frequency $\nu$) and that the cavity fields can be adiabatically eliminated from the system dynamics. In the approach we are following in this section, this leads to a reduced master equation for the density operator of the motion of the two atoms, given formally by \cite{Gardiner91} \begin{equation} \dot{\rho}_{\rm m} = {\rm Tr}_{\rm c} \left\{ {\cal L}_0 \int_0^\infty d\tau\; e^{{\cal L}_{\rm c}\tau} {\cal L}_0 \rho_{\rm c}^{\rm ss} \,\right\} \rho_{\rm m} , \end{equation} where $\rho_{\rm c}^{\rm ss}$ is the steady state density matrix for the two cavity modes. To evaluate this expression explicitly requires steady state correlation functions for the operators $\tilde{a}_1$ and $\tilde{a}_2$, which can be derived from (\ref{Lc}) using the quantum regression theorem. As shown in the appendix, the only nonzero correlation functions are \begin{equation} \label{corr1} \langle \tilde{a}_1(\tau )\tilde{a}_1^\dagger (0)\rangle_{\rm ss} = \langle \tilde{a}_2(\tau )\tilde{a}_2^\dagger (0)\rangle_{\rm ss} = e^{-\kappa\tau} \end{equation} and \begin{equation} \label{corr2} \langle \tilde{a}_1(0)\tilde{a}_2^\dagger (\tau )\rangle_{\rm ss} = \langle \tilde{a}_2(\tau )\tilde{a}_1^\dagger (0)\rangle_{\rm ss} = -2\kappa\tau e^{-\kappa\tau} \; . \end{equation} Using these expressions gives \begin{eqnarray} \dot{\rho}_{\rm m} &=& \Gamma_1(t) \left( 2\tilde{b}_1\rho_{\rm m}\tilde{b}_1^\dagger - \tilde{b}_1^\dagger\tilde{b}_1\rho_{\rm m} - \rho_{\rm m}\tilde{b}_1^\dagger\tilde{b}_1 \right) \nonumber \\ && + \Gamma_2(t) \left( 2\tilde{b}_2\rho_{\rm m}\tilde{b}_2^\dagger - \tilde{b}_2^\dagger\tilde{b}_2\rho_{\rm m} - \rho_{\rm m}\tilde{b}_2^\dagger\tilde{b}_2 \right) \nonumber \\ && + 2\sqrt{\Gamma_1(t)\Gamma_2(t)} \left\{ [\tilde{b}_2^\dagger ,\tilde{b}_1\rho_{\rm m} ] e^{-i(\phi_1-\phi_2)} \right. \nonumber \\ && \;\;\;\;\;\;\;\;\;\; \left. + [\rho_{\rm m}\tilde{b}_1^\dagger ,\tilde{b}_2 ] e^{i(\phi_1-\phi_2)} \right\} \, . \end{eqnarray} This master equation once again describes a cascaded system, only now the ``coupling'' appears directly between the motional modes of the two atoms. \subsection{ Quantum trajectories and ideal state transmission } To demonstrate the transmission properties of the coupled system, we again follow Cirac {\em et al.} and employ the technique of quantum trajectories \cite{Carmichael93,Zoller95}. This technique simulates a given master equation by propagating a system wave function $|\psi (t)\rangle$ subject to a non-Hermitian effective Hamiltonian. This propagation is interrupted at random times $\{ t_r\}$ by wave function collapses, or quantum jumps, $|\psi (t_r+dt)\rangle =\tilde{C}|\psi (t)\rangle$, which can be interpreted as, in our particular instance, the emission and destructive detection of photons from the cavity fields. For the master equation derived above, the effective Hamiltonian takes the form (choosing the laser phases such that $\phi_1=\phi_2$) \begin{eqnarray} \label{Heff} \tilde{H}_{\rm eff}(t) &=& -i\Gamma_1(t)\tilde{b}_1^\dagger \tilde{b}_1 -i\Gamma_2(t)\tilde{b}_2^\dagger \tilde{b}_2 \nonumber \\ && \;\;\; + 2i\sqrt{\Gamma_1(t)\Gamma_2(t)} \, \tilde{b}_2^\dagger \tilde{b}_1 \, , \end{eqnarray} while the collapse operator is given by \begin{equation} \tilde{C} = \sqrt{\Gamma_1(t)}\, \tilde{b}_1 - \sqrt{\Gamma_2(t)}\, \tilde{b}_2 \, . \end{equation} The basic idea is to design laser pulse profiles [manifest through $\Gamma_i(t)$] at the two sites such that the ideal quantum transmission \begin{equation} \label{perftrans} \sum_{n=0}^\infty c_n|n\rangle_1 \otimes |0\rangle_2 \rightarrow |0\rangle_1 \otimes \sum_{n=0}^\infty c_n|n\rangle_2 , \end{equation} can be achieved. Here $|n\rangle_i$ denotes the $n$-th Fock state of the motion of atom $i$. A necessary condition for successful transmission is that a quantum jump never occurs, i.e., $\tilde{C}|\psi (t)\rangle =0$ for all $t$, in which case the effective Hamiltonian becomes a Hermitian operator. This can be interpreted in terms of transfer via a {\em dark} state of the cascaded system. We expand the state of the system as \begin{equation} \label{psi_t} |\psi (t)\rangle = \sum_{n=0}^\infty c_n \sum_{m=0}^n \alpha_m^{(n)}(t) |n-m\rangle_1 \otimes |m\rangle_2 , \end{equation} with initial condition \begin{equation} \alpha_0^{(n)}(-\infty ) = 1 \, , \;\;\;\; \alpha_{m\neq 0}^{(n)}(-\infty ) = 0 , \end{equation} and normalization \begin{equation} \sum_{m=0}^n \left| \alpha_m^{(n)}(t) \right|^2 = 1 . \end{equation} For ideal quantum transmission, one requires that \begin{equation} \alpha_n^{(n)}(+\infty ) = \alpha_0^{(n)}(-\infty ) = 1 . \end{equation} Equations of motion for the $\{\alpha_m^{(n)}(t)\}$ are derived using (\ref{Heff}) and (\ref{psi_t}); the equations for the $m=0$ components take the simple closed form \begin{equation} \dot{\alpha}_0^{(n)}(t) = -n\Gamma_1(t)\alpha_0^{(n)}(t) , \end{equation} with solutions \begin{equation} \alpha_0^{(n)}(t) = \exp \left\{ -n\int_{-\infty}^t \Gamma_1(t^\prime ) dt^\prime \right\} \; . \end{equation} Now, applying the dark state condition \begin{equation} \left\{ \sqrt{\Gamma_1(t)}\, \tilde{b}_1 - \sqrt{\Gamma_2(t)}\, \tilde{b}_2 \right\} |\psi (t)\rangle = 0 \end{equation} yields the sequence of straightforward algebraic equations \begin{eqnarray} \sqrt{n\Gamma_1(t)}\,\alpha_0^{(n)}(t) - \sqrt{\Gamma_2(t)}\, \alpha_1^{(n)}(t) &=& 0 \, , \nonumber \\ \sqrt{(n-1)\Gamma_1(t)}\,\alpha_1^{(n)}(t) -\sqrt{2\Gamma_2(t)}\, \alpha_2^{(n)}(t) &=& 0 \, , \nonumber \\ \sqrt{(n-2)\Gamma_1(t)}\,\alpha_2^{(n)}(t) -\sqrt{3\Gamma_2(t)}\, \alpha_3^{(n)}(t) &=& 0 \, , \;\;\; \ldots \; , \end{eqnarray} from which a solution for $\alpha_m^{(n)}(t)$ in terms of $\alpha_0^{(n)}(t)$ follows as \begin{equation} \alpha_m^{(n)}(t) = \left[ \frac{\Gamma_1(t)}{\Gamma_2(t)} \right]^{m/2} \sqrt{\frac{n!}{m!(n-m)!}} \, \alpha_0^{(n)}(t) . \end{equation} Applying the normalization condition gives \begin{equation} \sum_{m=0}^n \left[ \alpha_m^{(n)}(t) \right]^2 = \left[ 1 + \frac{\Gamma_1(t)}{\Gamma_2(t)} \right]^n \left[ \alpha_0^{(n)}(t) \right]^2 = 1 \end{equation} and thus \begin{equation} \label{Gam12} \frac{\Gamma_1(t)}{\Gamma_2(t)} = \exp \left\{ 2\int_{-\infty}^t \Gamma_1(t^\prime ) dt^\prime \right\} - 1 , \end{equation} which demonstrates that it is possible to choose pulse shapes of the laser fields in such a way that the perfect transmission (\ref{perftrans}) can be achieved. We have not explored in detail pulse shapes satisfying (\ref{Gam12}). We do find, however, that (\ref{Gam12}) admits the following simple {\em analytical} solutions, \begin{equation} \Gamma_1(t) = \Gamma \, \frac{e^{\Gamma t}}{e^{\Gamma t}+e^{-\Gamma t}} \, , \;\;\;\; \Gamma_2(t) = \Gamma_1(-t) \, , \end{equation} with the limits $\Gamma_1(t)\rightarrow 0$ ($\Gamma$) as $t\rightarrow -\infty\; (+\infty )$, and {\em vice-versa} for $\Gamma_2(t)$. We use these forms in the numerical calculations that follow. \subsection{ Numerical calculations } For the purpose of numerical calculations, we retain the dynamics of the cavity field modes and solve the cascaded systems master equation \begin{eqnarray} \dot{\rho} &=& - \frac{i}{\hbar} [\tilde{H}_1(t)+\tilde{H}_2(t), \rho ] \nonumber \\ && +\, \kappa \left( 2\tilde{a}_1\rho\tilde{a}_1^\dagger - \tilde{a}_1^\dagger\tilde{a}_1\rho - \rho\tilde{a}_1^\dagger\tilde{a}_1 \right) \nonumber \\ && +\, \kappa \left( 2\tilde{a}_2\rho\tilde{a}_2^\dagger - \tilde{a}_2^\dagger\tilde{a}_2\rho - \rho\tilde{a}_2^\dagger\tilde{a}_2 \right) \nonumber \\ && -\, 2\kappa \, \left( [\tilde{a}_2^\dagger , \tilde{a}_1\rho ] + [\rho\tilde{a}_1^\dagger ,\tilde{a}_2] \right) \end{eqnarray} with \begin{eqnarray} \tilde{H}_k(t) &=& \hbar \Omega_k(t) \left( \tilde{a}_k^\dagger \tilde{b}_k e^{-i\phi_k} + \tilde{b}_k^\dagger \tilde{a}_k e^{i\phi_k} \right) \nonumber \\ &+& \, \hbar \Omega_k(t) \left( \tilde{a}_k^\dagger \tilde{b}_k^\dagger e^{-i\phi_k+2i\nu t} + \tilde{a}_k \tilde{b}_k e^{i\phi_k-2i\nu t} \right) \, , \end{eqnarray} where $k=1,2$. Once again, we include the rotating terms $\tilde{a}_k\tilde{b}_k$ and $\tilde{a}_k^\dagger\tilde{b}_k^\dagger$, while the forms of the time-dependent effective coupling parameters $\Omega_1(t)$ and $\Omega_2(t)$ are chosen in accordance with the work of the previous section, i.e., \begin{equation} \label{Om_overdamped} \Omega_1(t) = \Omega \, \sqrt{\frac{e^{\Gamma t}}{e^{\Gamma t}+e^{-\Gamma t}}} = \Omega_2(-t) \, , \end{equation} where $\Gamma = \Omega^2/\kappa$ (and we chosen $\phi_1=\phi_2=0$). As the state to be transferred, we choose, arbitrarily (in practice we are somewhat limited by the size of the basis set we can use comfortably in our simulations), \begin{equation} \label{psi} |\psi\rangle = \frac{1}{2} \left( |0\rangle + e^{i\pi /3}|1\rangle + e^{i2\pi /3}|2\rangle + e^{i\pi}|3\rangle \right) \, , \end{equation} so that the initial state of the total system is $|0\rangle_{a1} \otimes |\psi\rangle_{b1} \otimes |0\rangle_{a2} \otimes |0\rangle_{b2}$, while the target state is \begin{equation} |\psi_{\rm target}\rangle = |0\rangle_{a1} \otimes |0\rangle_{b1} \otimes |0\rangle_{a2} \otimes |\psi\rangle_{b2} \, . \end{equation} The transmission fidelity, which we define by \begin{equation} F(t) = \langle\psi_{\rm target}|\rho (t)|\psi_{\rm target}\rangle \, , \end{equation} is plotted in Fig.~5 for $\Omega /\kappa =0.141$ (corresponding to $\Gamma /\kappa =0.02$) and three different values of the trap frequency $\nu$. Note that the initial value of the fidelity is finite due to the contribution from $|0\rangle$ in the state $|\psi\rangle$. For $\nu /\kappa =20$ ($10$) the state is transmitted with a fidelity of 0.995 (0.980). As $\nu /\kappa$ is lowered the fidelity is reduced as the rotating terms in the effective Hamiltonians $\tilde{H}_k(t)$ begin to contribute more strongly to the dynamics. Nevertheless, a fidelity of 0.925 is found even for $\nu /\kappa =5$. In Fig.~6 we consider a single value of the trap frequency, $\nu /\kappa =20$, but now vary the coupling parameter $\Omega$, which varies the effective rate $\Gamma$ of the state transfer operation. We also choose larger values of $\Omega$ compared to $\kappa$ to assess how well the chosen pulse shapes drive the state transfer as the adiabatic approximation ($\kappa\gg\Omega$) ceases to be valid. As one can see, the fidelity of the transmission remains high even with $\Omega /\kappa =0.4$ (0.984) and $0.5$ (0.970), but at $\Omega /\kappa =0.7$ deteriorates to 0.905. Notably, the timescales for the transfer are significantly faster than in the previous figure, which is possibly advantageous from an experimental point of view. \subsection{ Underdamped regime } The overdamped regime considered above probably corresponds to the most likely experimental scenario. However, a situation where $\Omega >\kappa$ is still possible and could offer some further advantages in terms of transfer rates, which, in such a regime, would be of the order of $\kappa$. Of course, if $\Omega >\kappa$ then one must include the cavity modes and their dynamics in the analysis, which thereby becomes somewhat more complicated. Nevertheless, we find that it is still possible to derive pulse shapes that allow high-fidelity transmission of arbitrary quantum states. In particular, following the kind of approach used in \cite{Cirac97}, whereby one assumes that $\Omega_1(t)=\Omega$ (a constant) for $t\geq 0$ and that $\Omega_2(t)=\Omega_1(-t)$ (i.e., the {\em symmetric pulse condition}), we are able to arrive at the form \begin{equation} \label{eq:Om_underdamped} \Omega_1(-t) = -\,\frac{\Omega f(t)+2\kappa h(t)} {\sqrt{1-f(t)^2-2h(t)^2}} \;\;\;\; (t\geq 0), \end{equation} where \begin{eqnarray} f(t) &=& \frac{1}{p} \left( \lambda_+ e^{\lambda_-t} - \lambda_- e^{\lambda_+t} \right) f(0) \nonumber \\ && \;\;\;\;\; + \frac{\Omega}{p} \left( e^{\lambda_+t} - e^{\lambda_-t} \right) h(0) \, , \\ h(t) &=& \frac{1}{p} \left( \lambda_+ e^{\lambda_+t} - \lambda_- e^{\lambda_-t} \right) h(0) \nonumber \\ && \;\;\;\;\; - \frac{\Omega}{p} \left( e^{\lambda_+t} - e^{\lambda_-t} \right) f(0) \, , \end{eqnarray} with $\lambda_\pm = -(1/2)(\kappa \pm p)$, $p = (\kappa^2-4\Omega^2)^{1/2}$, and \begin{equation} f(0) = \sqrt{\frac{\kappa^2/2}{\kappa^2+\Omega^2}} \, , \;\;\;\; h(0) = -\sqrt{\frac{\Omega^2/2}{\kappa^2+\Omega^2}} \, . \end{equation} That such forms for the pulse shapes can successfully facilitate state transmission between the two atoms is illustrated by the dotted line in Fig.~6 for the same state $|\psi\rangle$ considered above and with $\Omega /\kappa =0.7$. The state is transmitted with a fidelity of 0.993, clearly improving on the result from the approach in the overdamped regime and also doing so on a faster timescale. \section{ Discussion } In this paper we have described schemes for the transfer of quite general quantum states between light fields and atomic motion and between the motion of trapped atoms at separate sites. We want now to consider in more detail some of the basic assumptions involved in our model and to examine possible experimental situations. Clearly, a very important assumption is that the effects of atomic spontaneous emission can be neglected. In a master equation approach, atomic spontaneous emission with the effects of recoil taken into account is modelled by a term of the form (considering motion only along the $x$ axis) \cite{Cirac92} \begin{equation} \{\dot{\rho}\}_{\rm spon} = \frac{\gamma}{2} \left( 2\hat{\sigma}_-\tilde{\rho}\hat{\sigma}_+ - \hat{\sigma}_+\hat{\sigma}_-\rho - \rho\hat{\sigma}_+\hat{\sigma}_- \right) \, , \end{equation} where \begin{eqnarray} \tilde{\rho} &=& \frac{1}{2} \int_{-1}^{+1} du \; W(u) e^{iku\hat{x}} \rho e^{-iku\hat{x}} \nonumber \\ &=& \frac{1}{2} \int_{-1}^{+1} du \; W(u) e^{i\eta_x (\hat{b}_x+\hat{b}_x^\dagger )} \rho e^{-i\eta_x (\hat{b}_x+\hat{b}_x^\dagger )} \, . \end{eqnarray} Here, $\gamma$ is the spontaneous emission rate and $W(u)=(3/4)(1+u^2)$ describes the angular distribution of spontaneous emission for an atomic dipole transition. Incorporating this into our analysis and adiabatically eliminating the atomic and cavity degrees of freedom as before, one finds that the leading order (in $\eta_x$) contribution to the motional dynamics contributed by atomic recoil due to spontaneous emission takes the form \begin{eqnarray} \eta_x^2 \, \frac{\gamma}{10} \frac{{\cal E}_{\rm L}^2}{\Delta^2} && \left[ 2(\hat{b}_x+\hat{b}_x^\dagger )\rho (\hat{b}_x+\hat{b}_x^\dagger ) \right. \nonumber \\ && \;\;\; - \left. (\hat{b}_x+\hat{b}_x^\dagger )^2\rho - \rho (\hat{b}_x+\hat{b}_x^\dagger )^2 \right] \, . \end{eqnarray} [Note that we also assume that ${\cal E}_{\rm L}\gg\eta_x g_0[\langle\hat{a}^\dagger\hat{a} \rangle ]^{1/2}$.] From inspection of the plots of transmission fidelity versus time (for the {\em overdamped} regime), we can estimate the timescale for state transfer as $\sim 4/\Gamma$. Hence, in order to be able to neglect the effects of spontaneous emission on the transfer process, we require that \begin{equation} \label{SpEcondition} \frac{\Gamma}{4} = \frac{\eta_x^2g_0^2{\cal E}_{\rm L}^2}{4\kappa\Delta^2} \gg \eta_x^2 \, \frac{\gamma}{10} \frac{{\cal E}_{\rm L}^2}{\Delta^2} \;\;\;\; {\rm or} \;\;\;\; \frac{5g_0^2}{2\kappa\gamma} \gg 1 \, . \end{equation} This, not surprisingly, corresponds to the regime of strong-coupling in cavity QED \cite{Kimble94}. The dipole coupling strength is given by $g_0 = [3c\lambda^2\gamma /(8\pi V_{\rm m})]^{1/2}$, where $\lambda$ is the wavelength of the atomic transition and $V_{\rm m}=(\pi /4)w_0^2l$ is the cavity mode volume, with $w_0$ the cavity mode waist and $l$ the mirror separation. The cavity field decay rate can be expressed in terms of the mirror separation and the cavity finesse ${\cal F}$ as $\kappa = \pi c/(2{\cal F}l)$. Using these expressions for $g_0$ and $\kappa$, the condition in (\ref{SpEcondition}) can be rewritten in the form \begin{equation} \frac{15}{2} \, \frac{\lambda^2}{\pi^3} \, \frac{{\cal F}}{w_0^2} \gg 1 . \end{equation} So, of course, one would like to have small cavity modes and high finesse mirrors. The trap itself must also meet rather stringent requirements; in particular, the Lamb-Dicke parameter must satisfy $\eta_j\ll 1$ ($j=x,y,z$) and the trap frequency along the $x$ axis must satisfy $\nu_x\gg\kappa$. Let us now consider a specific example from the ion-trapping community: the trapped ion species ${}^9{\rm Be}^+$. Recent experiments with this particular ion (see, for example, \cite{Meekhof96,Monroe96}) have been performed with harmonic oscillation frequencies along the principal axes of the trap $\nu_j/2\pi\simeq 11-30\;{\rm MHz}$, corresponding to Lamb-Dicke parameters $\eta_j\simeq 0.14-0.086$ (with respect to the $\,^2S_{1/2}\leftrightarrow\,^2P_{1/2}$ transition at wavelength $\lambda =313\;{\rm nm}$; the linewidth for this transition is $\gamma /2\pi =19.4\;{\rm MHz}$). A further practical consideration is the requirement that the spacing between the mirrors be large enough to accommodate the ion trap electrodes and external laser fields. A reasonable minimum separation might be $l=100\;\mu {\rm m}$, but we shall make a somewhat more conservative choice of $l=250\;\mu {\rm m}$. With a cavity finesse ${\cal F}=300,000$ \cite{Mirrors}, one then obtains $\kappa /2\pi =1.0\;{\rm MHz}$. Assuming, say, that $\eta_x=0.1$, then $\nu_x/2\pi =22\;{\rm MHz}$ and $\nu_x/\kappa =22$. If the radius of curvature of the mirrors is taken to be 5 cm, then the cavity waist takes the value $w_0=15.8\;\mu {\rm m}$ (which yields $g_0/2\pi =14.9\;{\rm MHz}$) and $15\lambda^2{\cal F}/(2\pi^3w_0^2)=28\gg 1$. Finally, given these choices of parameters, a numerical estimate for the rate $\Gamma =\Omega^2/\kappa$ at which the state transfer occurs in the overdamped regime would be $\Gamma /2\pi\simeq 20-200\;{\rm kHz}$, and even larger with suitable laser pulse shapes in the underdamped regime. This estimate establishes a timescale for the required stability of the driving laser fields (remember that the phases of the two laser fields are assumed to be equal for the duration of the state transfer process) and of the cavity and trap setups. Note that the timescales for motional decoherence and heating observed in recent trapped ion experiments (with ${}^9{\rm Be}^+$) are of the order of milliseconds and further improvement seems possible \cite{Wineland98}. So, it would seem that, with quite reasonable choices of experimental parameters, a suitable operating regime for the state transfer scheme is feasible. Note that favorable parameters should also be achievable with trapped-ion species other than ${}^9{\rm Be}^+$; for example, ${}^{24}{\rm Mg}^+$ or ${}^{40}{\rm Ca}^+$. Alternatively, recent developments with microscopic magnetic traps \cite{Weinstein95,Vuletic98} suggest that suitably large trap frequencies and confinement in the Lamb-Dicke regime may also be possible with neutral atoms, such as Li and K. To conclude, in this work we have described a means of usefully combining several emerging candidate technologies for the implementation of quantum communication and computing, i.e., trapped atoms, cavity QED, and propagating (nonclassical) light fields. The schemes outlined above allow, in principle, for the transfer of quantum states and entanglement between light fields and motional degrees of freedom of trapped atoms and for high-fidelity transmission of quantum states between spatially distant sites. \acknowledgments We gratefully acknowledge the contributions of Prof. S. Braunstein, whose insight was crucial in initiating the present work. ASP thanks A. Doherty, C. Hood, Q. Turchette, S. van Enk, L. You and P. Zoller for helpful discussions and the ITP in Santa Barbara for its hospitality, where part of this work was carried out, supported in part by the NSF under Grant No. PHY94-07194. ASP also thanks the Quantum Optics Group at Caltech for its hospitality and acknowledges support from the Marsden Fund of the Royal Society of New Zealand. HJK is supported by the National Science Foundation, by DARPA via the QUIC Institute which is administered by ARO, and by the Office of Naval Research.
\section{Introduction} \end{center} \renewcommand{\theequation}{1.\arabic{equation}}\setcounter{equation}{0} A class of developments of quantum field theory in the ninety of this century may be represented by two keywords: $N=2$ supersymmetry and duality. For example, the mirror symmetry \cite{Yau,M1,Can,Can2} established in the beginning of the ninety was based on the (trivial) isomorphism between left and right $U(1)$ currents of $(2,2)$-superconformal field theory,\cite{Dix,GP2} and this isomorphism predicted the existence of a pair of Calabi-Yau manifolds whose axes of Hodge diamond were exchanged. Candelas {\em et al.} \cite{Can,Can2} skillfully used the consequence expected from this duality of Hodge structure and showed that the numbers of rational curves on Calabi-Yau quintic 3-fold could be determined from mirror symmetry. The coincidence of their result with mathematically rigorous results \cite{Yau,M2,Kat} gave a great surprising! On the other hand, also in the recent studies of low energy effective dynamics of $N=2$ supersymmetric Yang-Mills theory, $N=2$ supersymmetry and duality play a crucial role. Before the arrival of Seiberg and Witten's proposal by using electro-magnetic duality for the description of the low energy effective action of SU(2) gauge theory, \cite{SW1,SW2} though it has been known that the prepotential which is a generating function of the low energy effective action is not renormalized beyond one-loop in perturbative calculation due to $N=2$ supersymmetry, \cite{HST1,HST2,HSW} actually the prepotential was expected to receive instanton corrections. \cite{Sei} Unfortunately, such corrections were not so extensively discussed, but thanks to their proposal, it made possible to extract informations on instanton effects from a Riemann surface and periods of meromorphic differential on it. Namely, the low energy effective theory was turned out to be parameterized by a Riemann surface. The validity of their proposal was discussed by Klemm {\em et al.} \cite{KLT} with the aid of Picard-Fuchs equation and the instanton corrections to the prepotential was revealed. The instanton corrections obtained in this way showed extremely good agreement with the prediction of instanton calculus. \cite{FP,IS3,IS,Slat,DKM1,AHSW} However, deeper and striking features of prepotentials of $N=2$ supersymmetric Yang-Mills theories may be nicely interpreted in terms of differential equations satisfied by prepotentials. For instance, it is well-known that the prepotentials satisfy an Euler equation called scaling relation, \cite{Mat,STY,EY,HW,KO} and in fact this simple equation simplified and accelerated the study of prepotentials. As for another characteristic equations, we can mention that there is a non-linear system of partial differential equations called Witten-Dijkgraaf-Verlinde-Verlinde (WDVV) equations \cite{MMM1,MMM2,MMM3,IY,IXY} (rigorously speaking, the WDVV equations in $N=2$ Yang-Mills theory are not equivalent to those arising in two-dimensional topological field theory \cite{Wit,DW,DVV,Dub}). Actually, these equations hold not only in four-dimensional gauge theories but also in higher dimensions even if hypermultiplets are included. \cite{MMM2,MMM3} Accordingly, it becomes possible to regard the prepotentials in various gauge theories as a member of solutions to WDVV equations. Then, what is the most general solution (function form of prepotentials) to the WDVV equations? Unfortunately, we can not precisely know the answer to this question, but Braden {\em et al.} \cite{BMMM} partially found the answer. They assumed the function form of prepotential which is expected from known examples and found a new prepotential which is considered as that in five-dimensional gauge theory, although their study was restricted to perturbative part. Of course, among the solutions found by them we can see the existence of the prepotential in four-dimensional Yang-Mills theory. This seems to indicate that the prepotentials can be constructed without introducing Riemann surface, provided the WDVV equations are used. Finding whether non-perturbative prepotentials are available from WDVV equations without using Riemann surface is the subject of this paper. The paper is organized as follows. In Sec. II, the construction of perturbative solution to WDVV equations for SU(4) gauge theory in four dimensions discussed by Braden {\em et al.} \cite{BMMM} is summarized. We can see that the perturbative prepotential is in fact obtained from WDVV equations. In Sec. III, we add the non-perturbative part for this perturbative prepotential and try to solve the WDVV equations. Though the non-perturbative part satisfies a non-linear differential equation, restricting it at one-instanton level, we can reduce it to a linear differential equation satisfied by one-instanton prepotential. For this reason, the one-instanton prepotential is investigated in this paper. To solve this equation, the scaling relation is used as a subsidiary condition, but it turns out that there are miscellaneous solutions which do not contradict to both WDVV equations and scaling relation. In Sec. IV. we compare our result with the prediction of one-instanton calculus. It is shown that among our one-instanton prepotentials obtained from WDVV equations there are one-instanton prepotentials which agree the prediction of instanton calculus. In this way, we conclude that it is possible to obtain non-perturbative prepotential from WDVV equations without relying on Riemann surface. Sec. V is a brief summary. \begin{center} \section{Perturbative prepotential from WDVV equations } \end{center} \renewcommand{\theequation}{2.\arabic{equation}}\setcounter{equation}{0} In this section, we briefly outline the construction to get perturbative prepotential from WDVV equations in the SU(4) gauge theory presented by Braden {\em et al}. \cite{BMMM} Note that the SU(4) model is the simplest and non-trivial example for a study of WDVV equations. In this case, the WDVV equations for the prepotential ${\cal{F}}$ take the form \beq ({\cal{F}}_i )({\cal{F}}_k )^{-1}({\cal{F}}_j )=({\cal{F}}_j )({\cal{F}}_k )^{-1}({\cal{F}}_i ) ,\ i,j,k =1,\cdots,3 ,\lab{wdvv} \end{equation} where \beq ({\cal{F}}_i )\equiv ({\cal{F}}_i )_{jk}=\frac{\partial^3 {\cal{F}}}{\partial a_i \partial a_j \partial a_k} \end{equation} are the matrix notations and in this paper the brackets are always added as $({\cal{F}}_i )$ when ``${\cal{F}}_i$'' mean matrices. The coordinates $a_i$ are the periods of the SU(4) gauge theory (see Appendix). Braden {\em et al.} \cite{BMMM} considered the perturbative prepotential in the form \beq {\cal{F}}_{\mbox{\scriptsize per}}(a_1 ,a_2 ,a_3 ) = \sum_{i<j=1}^{4}f(a_{ij}),\ a_{ij}=a_i -a_j , \ \sum_{i=1}^4 a_i =0 .\lab{22} \end{equation} Of course ${\cal{F}}_{\mbox{\scriptsize per}}$ may depend on the mass scale $\Lambda \equiv \Lambda_{\mbox{\scriptsize SU(4)}}^8$ of the theory, but we can ignore its dependence for the moment because $\Lambda$-differentiation is not included in the WDVV equations (\ref{wdvv}). Under the assumption (\ref{22}), we can find that when ${\cal{F}}_{\mbox{\scriptsize per}}$ satisfies (\ref{wdvv}) there is a functional relation \beq g(a_{12})g(a_{34})-g(a_{13})g(a_{24})+g(a_{14})g(a_{23})=0 ,\lab{23} \end{equation} where \beq g(a)\equiv \left(\frac{\partial^3 f}{\partial a^3}\right)^{-1} .\lab{ga} \end{equation} With the aid of several conditions, we can conclude that $g$ is an odd function with \beq g(0)=g^{\,''}(0)=0, \lab{226} \end{equation} where the prime means the differentiation over the argument. \cite{BMMM} There are several functions enjoying the properties (\ref{23}) and (\ref{226}), but a function which is necessary for us among them is the function of the form $g(a)=a$. Namely, \beq f(a)=\frac{a^2}{2}\ln a+O(a^2) .\lab{26} \end{equation} Note that $O(a^2)$-term can not be fixed from the WDVV equations because they are third-order differential equations. Namely, $\Lambda$-dependence of one-loop contribution is not fixed. It is easy to see that substituting (\ref{26}) back to (\ref{22}) in fact yields the perturbative part of the SU(4) prepotential (in a suitable normalization). Marshakov {\em et al.} \cite{MMM3} give a general proof that the perturbative prepotentials in various gauge theories satisfy WDVV equations, but this is also confirmed by Ito and Yang \cite{IY} in their study of these equations. \begin{center} \section{One-instanton prepotentials} \end{center} \renewcommand{\theequation}{3.\arabic{equation}}\setcounter{equation}{0} \begin{center} \subsection{Differential equation for one-instanton prepotential} \end{center} Next, let us consider whether non-perturbative prepotential ${\cal{F}}$ is available from the WDVV equations by assuming the form \beq {\cal{F}} (a_1 ,a_2 ,a_3 ,\Lambda )={\cal{F}}_{\mbox{\scriptsize per}}(a_1 ,a_2 ,a_3 ) +{\cal{F}}_{\mbox{\scriptsize ins}}(a_1 ,a_2 ,a_3 ,\Lambda ) ,\lab{2} \end{equation} where \beq {\cal{F}}_{\mbox{\scriptsize ins}}(a_1 ,a_2 ,a_3 ,\Lambda )= \sum_{k=1}^{\infty}{\cal{F}}_k (a_1 ,a_2 ,a_3 ) \Lambda^k .\end{equation} In order to derive differential equations for ${\cal{F}}_{\mbox{\scriptsize ins}}$, we assume that ${\cal{F}}_{\mbox{\scriptsize per}}$ is already given by (\ref{22}) with (\ref{26}). Then substituting ${\cal{F}}$ into (\ref{wdvv}) we can obtain a single non-linear differential equation for ${\cal{F}}_{\mbox{\scriptsize ins}}$, but if we restrict only the case $k=1$ (one-instanton level), the equation reduces to \begin{eqnarray} & &\frac{\partial_{1}^3 {\cal{F}}_1}{a_{12}a_{13}a_{14}}-\frac{\partial_{2}^3 {\cal{F}}_1} {a_{12}a_{23}a_{24}}+\frac{\partial_{3}^3 {\cal{F}}_1}{a_{13}a_{23}a_{34}} +\frac{6\partial_1\partial_2\partial_3{\cal{F}}_1}{a_{14}a_{24}a_{34}} -\frac{A_{012}\partial_2\partial_{3}^2{\cal{F}}_1 +A_{021}\partial_{2}^2\partial_3{\cal{F}}_1} {a_{12}a_{13}a_{23}a_{24}a_{34}}\nonumber\\ & &-\frac{A_{102}\partial_1\partial_{3}^2 {\cal{F}}_1 -A_{201}\partial_{1}^2\partial_3 {\cal{F}}_1}{a_{12}a_{13}a_{14}a_{23}a_{34}} +\frac{A_{120}\partial_1\partial_{2}^2{\cal{F}}_1 +A_{210}\partial_{1}^2\partial_2{\cal{F}}_1} {a_{12}a_{13}a_{14}a_{23}a_{24}}=0 \lab{k} ,\end{eqnarray} where $\partial_i \equiv \partial /\partial a_i$ and \begin{eqnarray} & &A_{012}=a_1 a_2 -3a_{2}^2 -2a_1 a_3 +4a_2 a_3+a_1 a_4 +a_2 a_4 -2 a_3 a_4 ,\nonumber\\ & &A_{021}=2a_1 a_2 -a_1 a_3 -4a_2 a_3 +3a_{3}^2 -a_1 a_4 +2 a_2 a_4 -a_3 a_4 ,\nonumber\\ & &A_{102}=3a_{1}^2 -a_1 a_2 -4a_1 a_3 +2a_2 a_3 -a_1 a_4 -a_2 a_4 +2a_3 a_4 ,\nonumber\\ & &A_{201}=2a_1 a_2 -4a_1 a_3 -a_2 a_3 +3a_{3}^2 +2a_1 a_4 -a_2 a_4 -a_3 a_4 ,\nonumber\\ & &A_{120}=3a_{1}^2 -4a_1 a_2 -a_1 a_3 +2a_2 a_3 -a_1 a_4 +2a_2 a_4 -a_3 a_4 ,\nonumber\\ & &A_{210}=4a_1 a_2 -3a_{2}^2 -2a_1 a_3 +a_2 a_3 -2a_1 a_4 + a_2 a_4 +a_3 a_4 .\end{eqnarray} \begin{center} \subsection{The solutions} \end{center} In order to solve (\ref{k}), let us introduce the new variables \beq x=a_{12} ,\ y=a_{13},\ z=a_{14} \lab{111} .\end{equation} In addition, using Euler derivatives $\theta_{x} =x\partial /\partial x$ etc, we can rewrite (\ref{k}) as \beq L(\theta_{x} ,\theta_{y} ,\theta_{z} ){\cal{F}}_1 =0 ,\end{equation} where \begin{eqnarray} L(\theta_{x} ,\theta_{y} ,\theta_{z} )&=&yz(y-z)\theta_{x} (\theta_{x} -1)(\theta_{x} -2)+z(4xy-3y^2 -2xz+yz) (\theta_{x} -1)\theta_{x} \theta_{y} \nonumber\\ & &+z(3x^2 -4xy-xz +2yz )(\theta_{y} -1)\theta_{x} \theta_{y} -xz(x-z)\theta_{y} (\theta_{y}-1)(\theta_{y} -2) \nonumber\\ & &-y(3x^2 -xy-4xz +2yz)(\theta_{z} -1)\theta_{x} \theta_{y} -y(4xz+yz -3z^2 -2xy )(\theta_{x} -1)\theta_{x} \theta_{z} \nonumber\\ & &+6(x-y)(x-z)(y-z)\theta_{x} \theta_{y} \theta_{z} + x(xz+4yz -3z^2 -2xy)(\theta_{y} -1)\theta_{y} \theta_{z} \nonumber\\ & &+x(3y^2 +2xz -4yz -xy)(\theta_{z} -1)\theta_{y} \theta_{z} + xy(x-y)\theta_{z} (\theta_{z} -1)(\theta_{z} -2) .\end{eqnarray} Here, suppose that ${\cal{F}}_1$ is given by \beq {\cal{F}}_1 =x^{\nu_1}y^{\nu_2}z^{\nu_3}F(x,y,z) \lab{nu} ,\end{equation} where \beq F(x,y,z)= \sum_{i,j,k=0}^{\infty}B_{\epsilon_{x} i,\epsilon_{y} j,\epsilon_{z} k}x^{\epsilon_{x} i} y^{\epsilon_{y} j} z^{\epsilon_{z} k} \lab{b} \end{equation} and the expansion coefficients are assumed to be independent of $x,y$ and $z$. In (\ref{b}), $\epsilon_{x}$ etc are signature symbols (the choice of signatures depends on where is the convergence region), e.g., $\epsilon_{x} =\pm$. Then from (\ref{k}) we get the differential equation for $F$ \beq L (\theta_{x} +\nu_{1} ,\theta_{y} +\nu_{2} ,\theta_{z} +\nu_{3} )F =0 .\lab{cc} \end{equation} and the indicial equations for $\nu_i$ \begin{eqnarray} & &\nu_{1} (\nu_{1} -2\nu_{2} -1)(\nu_{1} +\nu_{2} -3\nu_{3} -2)=0,\ \nu_{2} (2\nu_{1} -\nu_{2} +1)(\nu_{1} +\nu_{2} -3\nu_{3} -2)=0,\nonumber\\ & &\nu_{2} (\nu_{2} -2\nu_{3} -1)(3\nu_{1} -\nu_{2} -\nu_{3} +2)=0,\ (\nu_{2} -\nu_{3} )(\nu_{1}^2 -\nu_{1} -\nu_{1}\nu_{2} +\nu_{2} \nu_{3} )=0,\nonumber\\ & &\nu_{1} (\nu_{1}^2 -3\nu_{1} +5\nu_{2} -3\nu_{1}\nu_{2} +\nu_{3} -\nu_{1}\nu_{3} +4\nu_{2}\nu_{3} +2)=0 ,\nonumber\\ & &\nu_{3}^2 (\nu_{3} -3\nu_{2}-3)+5\nu_{1}\nu_{2}\nu_{3} +3\nu_{2}\nu_{3} -2\nu_{1}^2 \nu_{3}+ 2\nu_{1}\nu_{3}+2\nu_{3} +\nu_{1}\nu_{2} =0,\nonumber\\ & &\nu_{3}^2 (\nu_{3} -\nu_{2}-3)-2\nu_{2}^2 \nu_{3} +3\nu_{1}\nu_{2}\nu_{3} +3\nu_{2}\nu_{3} +3\nu_{1}\nu_{2} + 2\nu_{3} =0 .\end{eqnarray} The sets of possible $\nu_i$ are thus given by \begin{eqnarray} \mbox{\boldmath$\nu$}\equiv (\nu_{1},\nu_{2},\nu_{3})&=&(-2,-2,-2),\ (-1,-1,-1),\ (0,-1,-1)^4 ,\ (0,0,0)^2 ,\ (0,0,1),\ (0,0,2)^2 ,\nonumber\\ & & (0,1,0),\ (0,1,1),\ (0,2,0)^2 ,\ (1,0,0)^2 ,\ (1,0,1),\ (1,0,2),\ (2,0,0)^2 ,\lab{list} \end{eqnarray} where the superscript means degeneracy, e.g., $(0,0,0)^2$ is composed of two $(0,0,0)$, but we do not discuss the consequence of degeneracy in this paper. For the indices (\ref{list}), it is straightforward to obtain $F$. Though we have expressed $F$ in (\ref{b}) as an infinite series, actually we can restrict possible terms in (\ref{b}) by considering the degree counting of one-instanton prepotential. \begin{center} \subsection{Scaling relation for one-instanton prepotential} \end{center} If the WDVV equations can in fact yield a physically acceptable prepotential, the prepotential obtained from those equations must also satisfy the fundamental homogeneity condition called scaling relation. \cite{Mat,STY,EY,HW,KO} Therefore, we may use it as a subsidiary condition for the problem how to solve WDVV equations in gauge theory. To see this, firstly, let us recall the scaling relation \cite{Mat,STY,EY,HW,KO} \beq \sum_{i=1}^{3}a_i \frac{\partial {\cal{F}}}{\partial a_i} + \Lambda_{\mbox{\scriptsize SU(4)}} \frac{\partial {\cal{F}}} {\partial \Lambda_{\mbox{\scriptsize SU(4)}}}=2{\cal{F}} .\lab{sca} \end{equation} We need a scaling relation for ${\cal{F}}_1$ not for ${\cal{F}}$ itself, but in order to extract it from (\ref{sca}), the $\Lambda$-dependence of perturbative prepotential which can not be fixed from WDVV equations must be included appropriately. For this, our choice here is \beq {\cal{F}}_{\mbox{\scriptsize per}}=\sum_{i<j=1}^{4}\frac{a_{ij}^2}{2} \ln \frac{a_{ij}}{\Lambda_{\mbox{\scriptsize SU(4)}}} .\end{equation} Then from (\ref{sca}), ${\cal{F}}_1$ is found to satisfy \beq \sum_{i=1}^3 a_i \frac{\partial {\cal{F}}_1}{\partial a_i} +6{\cal{F}}_1 =0 \lab{42} ,\end{equation} which indicates that ${\cal{F}}_1$ is a homogeneous function of degree $-6$. In the variables (\ref{111}), (\ref{42}) becomes \beq x\partial_x {\cal{F}}_1 +y\partial_y {\cal{F}}_1 +z\partial_z {\cal{F}}_1 +6{\cal{F}}_1 =0 .\lab{318} \end{equation} Accordingly, from (\ref{nu}), (\ref{b}) and (\ref{318}) it must be always true that \beq \nu_{1} +\nu_{2} +\nu_{3} +\epsilon_{x} i+\epsilon_{y} j+\epsilon_{z} k =-6 .\lab{319} \end{equation} \begin{center} \subsection{Examples of one-instanton prepotentials} \end{center} We have now enough informations to construct explicit one-instanton prepotentials which do not contradict to WDVV equations and scaling relation. To begin with, let us consider the case $\mbox{\boldmath$\epsilon$}\equiv (\epsilon_{x} ,\epsilon_{y} ,\epsilon_{z})=(+,+,+)$. In this case, we can easily find that there exists only one solution which satisfies (\ref{319}). It is the solution with $\mbox{\boldmath$\nu$}=(-2,-2,-2)$, and thus \beq {\cal{F}}_1 =\frac{B_{0,0,0}}{x^2 y^2 z^2 } .\lab{f1} \end{equation} However, when $\mbox{\boldmath$\epsilon$}=(-,-,-)$ or one entry of $\mbox{\boldmath$\epsilon$}$ differs to the others, e.g., $\mbox{\boldmath$\epsilon$} =(-,-,+)$, the situation changes, in particular, drastically in the latter case. In the former case, it is easy to see that $F$ consists of finite number of terms for all indices in (\ref{list}), but in the latter case $F$ is generally represented by infinite number of terms as long as (\ref{319}) is satisfied. We do not know whether it is possible to find any physical meaning for this type of one-instanton prepotential, but it may be interesting to recall that a similar one was observed in the one-instanton prepotential in the five-dimensional gauge theory.\cite{KO} Since the latter case mentioned above is slightly intractable, let us consider an example of the former case instead. In the case of $\mbox{\boldmath$\nu$}=(-1,-1,-1)$ with $\mbox{\boldmath$\epsilon$}=(-,-,-)$, for instance, we have \begin{eqnarray} {\cal{F}}_1 &=&\frac{1}{xyz} \left[B_{0,-1,-2}\left(\frac{1}{yz^2}+\frac{1}{y^2 z}\right) +B_{-1,0,-2}\left(\frac{1}{xz^2}+\frac{1}{x^2 z}\right)\right.\nonumber\\ & &\left. +B_{-1,-2,0}\left( \frac{1}{x^2 y}+\frac{1}{xy^2}\right) +B_{-1,-1,-1}\frac{1}{xyz} \right] .\lab{3333} \end{eqnarray} Note that (\ref{3333}) includes the one-instanton prepotential of the form (\ref{f1}). In a similar manner, we can construct one-instanton prepotentials for all other possible values of $\nu_i$, which do not contradict to both WDVV equations and scaling relation, but it would not be necessary to explicitly show them here. However, we should point out that since (\ref{k}) is a partial differential system we can expect that there exist more and more various solutions. In fact, this observation is right, and we can show that also in the variables \beq (x,y,z)=(a_{12},a_{23},a_{24}),\ (a_{13}, a_{23},a_{34}) \lab{lat} \end{equation} we can construct miscellaneous one-instanton prepotentials. Among them one-instanton prepotentials of the form (\ref{f1}) are included. {\bf Remark:} {\em The function form of the one-instanton prepotentials can be determined by solving the WDVV equations, but its numerical factors, i.e., instanton expansion coefficients, are not obtained because they correspond to integration constants. In order to determine them, it is necessary to rewrite the scaling relation as a relation between prepotential and moduli. Then substituting the one-instanton prepotential obtained from WDVV equations into this scaling relation, we will be able to get the expansion coefficients. Of course, in this case the moduli must be represented as a function of periods and its expansion coefficients must be determined. However, since knowing moduli is equivalent to introduce a Seiberg-Witten curve, this method based on scaling relation represented by using moduli is not preferable in the formalism of WDVV equations because prepotentials available from WDVV equations should be determined without introduction of Seiberg-Witten curves. Accordingly, when the determination of instanton expansion coefficients is required, they should be determined from the result of instanton calculus.} \begin{center} \section{One-instanton prepotential from instanton calculus} \end{center} \renewcommand{\theequation}{4.\arabic{equation}}\setcounter{equation}{0} We have derived one-instanton prepotentials by solving the WDVV equations in the previous section. Though these one-instanton prepotentials satisfies the WDVV equations and the scaling relation, unfortunately in view of WDVV equations we can not determine which ones are physically acceptable. For this reason, in order to extract physically meaningful one-instanton prepotentials among them, we must compare our result with the one-instanton prediction of instanton calculus. In the case of SU(4) gauge theory, one-instanton contribution for prepotential is given by \cite{IS3,IS} \beq {\cal{F}}_1 =\frac{\Delta_{4}^{\,'}}{\Delta_4} ,\lab{37} \end{equation} where we have omitted the numerical normalization factor and \beq \Delta_{4}^{\,'}=\sum_{i=1}^{4} \prod_{\stackrel{k<l=1}{k,l\neq i}}^{4}(a_k -a_l )^2 ,\ \Delta_4 =\prod_{k<l=1}^{4}(a_k -a_l )^2 .\end{equation} The closed form of one-instanton prepotential for SU($N_c$) gauge theory is also obtained by solving Picard-Fuchs equations \cite{IY2} and direct calculation of period integrals. \cite{DKP,MS} Note that (\ref{37}) is a sum of (\ref{f1}) and those for (\ref{lat}) \beq {\cal{F}}_1 =\frac{1}{(a_{12}a_{13}a_{14})^2}+\frac{1}{ (a_{12}a_{23}a_{24})^2}+\frac{1}{(a_{13}a_{23}a_{34})^2} \lab{310} \end{equation} up to constant factors. Accordingly, we can conclude that the WDVV equations can yield physical prepotential in spite of without introducing Riemann surface. \begin{center} \section{Summary} \end{center} \renewcommand{\theequation}{4.\arabic{equation}}\setcounter{equation}{0} In this paper, we have discussed the non-perturbative prepotential of $N=2$ supersymmetric SU(4) Yang-Mills theory in the standpoint of WDVV equations. Especially, we have found a differential equation for one-instanton prepotential and constructed its solutions. The method to get prepotentials based on WDVV equations is fascinating in the point that the prepotentials can be obtained without introducing Seiberg-Witten curves, but it has been shown that unfortunately too many prepotentials exist in contrast with the approach based on Seiberg-Witten curves which uniquely determines a prepotential. Nevertheless, we have succeeded to show that one-instanton prepotentials which coincide with the one-instanton calculus can be obtained from WDVV equations. As for an another aspect of WDVV equations, we should mention a connection to topological field theory in two dimensions. From the appearance of WDVV equations, it may be natural to think that the low energy effective theory is actually a kind of topological field theory, but we must notice that a priori there is no reason that the effective theory must be a topological field theory. Therefore, topological or not topological: that is the question. An approach to argue this implication more explicitly is to regard the Seiberg-Witten curves often identified with spectral curves of integrable system as if they were superpotentials of topological $\mbox{\boldmath$CP$}^1$ model. \cite{IXY} Although this observation strongly relying on the existence of Riemann surface (Seiberg-Witten curve), it enables us to find a connection to topological field theory, specifically, topological string theory at genus zero level. \cite{IXY} Then, even if we do not assume a Riemann surface, can we find a topological nature of the effective theory? Probably WDVV equations give the answer, but the study is a subject in the future. \begin{center} \section*{Acknowledgment} \end{center} The author acknowledges Prof. H. Kanno for discussions about WDVV equations. \begin{center} \section*{Appendix: The SU(4) Seiberg-Witten solution} \end{center} \renewcommand{\theequation}{A\arabic{equation}}\setcounter{equation}{0} In this appendix, we briefly summarize the SU(4) Seiberg-Witten solution. The SU(4) Seiberg-Witten curve is given by the hyperelliptic curve of genus three \cite{KLYT,AF,HO,AAM} \beq y^2 =(x^4 -ux^2 -vx-w)^2 -\Lambda_{\mbox{\scriptsize SU(4)}}^8 ,\end{equation} where $(x,y)\in \mbox{\boldmath$C$}^2$ is the local coordinate, $u,v$ and $w$ are moduli of the theory. Then the Seiberg-Witten differential and its periods are given by \beq \lambda_{\mbox{\scriptsize SW}}=\frac{x\partial_x W}{y}dx \end{equation} and \beq a_i =\oint_{\alpha_i}\lambda_{\mbox{\scriptsize SW}},\ a_{D_i}=\oint_{\beta_i}\lambda_{\mbox{\scriptsize SW}} ,\ i=1,2,3 ,\end{equation} respectively, where $\alpha_i$ and $\beta_i$ are the canonical bases of the 1-cycles on the curve and the numerical normalization factor of the Seiberg-Witten differential is ignored. It is convenient to use the period vector \beq \Pi =\left(\begin{array}{c} a_{D_i}\\ a_i \end{array}\right) .\end{equation} In general, these periods satisfy Fuchsian differential equations and in the case at hand they are given by \cite{KLT,IS,IMNS,Ali} \begin{eqnarray} {\cal{L}}_1 \Pi &\equiv& \left[ \partial_{v}^2 -\partial_u \partial_w\right]\Pi=0,\nonumber\\ {\cal{L}}_2 \Pi &\equiv& \left[ 4\partial_{u}^2 -2u\partial_u \partial_w - v\partial_v\partial_w -\partial_w \right]\Pi=0,\nonumber\\ {\cal{L}}_3 \Pi &\equiv& \left[ v\partial_{w}^2 +2u\partial_v\partial_w -4\partial_u\partial_v \right]\Pi=0,\nonumber\\ & &\left[ 4(u^2 +24w)\partial_{u}^2 +9v^2\partial_{v}^2 -16 (\Lambda_{\mbox{\scriptsize SU(4)}}^8 -w^2 )\partial_{w}^2 +12uv\partial_u\partial_v \right.\nonumber\\ & &\left.-32uw\partial_u\partial_w +3v\partial_v -16w\partial_w +1 \right]\Pi=0 ,\lab{imns1} \end{eqnarray} which can be summarized as \begin{eqnarray} & &\left[\theta_{v} (\theta_{v} -1) -\frac{v^2}{uw}\theta_{u}\theta_{w} \right]\Pi =0,\nonumber\\ & &\left[(2\theta_{u} +\theta_{v} +1)\theta_{w} -\frac{4w}{u^2}\theta_{u} (\theta_{u} -1) \right]\Pi =0,\nonumber\\ & &\left[(2\theta_{u} +3\theta_{v} +4\theta_{w} -1)^2 -\frac{16\Lambda_{\mbox{\scriptsize SU(4)}}^8}{w^2}\theta_{w} (\theta_{w} -1)\right]\Pi =0 ,\lab{ne} \end{eqnarray} where $\theta_{w} =w\partial_w$ etc are Euler derivatives, provided the first, second and last equations in (\ref{imns1}) are chosen as independent equations. Note that the third one in (\ref{imns1}) is not an independent equation because $(v\partial_w {\cal{L}}_1 +\partial_v {\cal{L}}_2 +\partial_u {\cal{L}}_3 )\Pi =0$. Introducing new variables $x,y$ and $z$ by \beq x=\frac{\Lambda_{\mbox{\scriptsize SU(4)}}^8}{4w^2},\ y=\frac{v^2}{4uw},\ z=\frac{w}{u^2} ,\end{equation} we find that (\ref{ne}) is converted into \begin{eqnarray} & &\left[(8\theta_{x} +1)^2 -64x(2\theta_{x}+\theta_{y}-\theta_{z})(2\theta_{x}+\theta_{y}-\theta_{z} +1) \right]\Pi =0,\nonumber\\ & &\left[\theta_{y} (2\theta_{y} -1) -2y(\theta_{y}+2\theta_{z}) (2\theta_{x}+\theta_{y} -\theta_{z})\right]\Pi =0,\nonumber\\ & &\left[(2\theta_{x}+\theta_{y}-\theta_{z})(4\theta_{z} -1)-4z (\theta_{y} +2\theta_{z})(\theta_{y}+2\theta_{z} +1) \right]\Pi =0 .\lab{SU4PF} \end{eqnarray} This system (\ref{SU4PF}) further simplifies to \begin{eqnarray} & &\left[\theta_{x}^2 -x\left(2\theta_{x}+\theta_{y}-\theta_{z}-\frac{1}{2}\right) \left(2\theta_{x}+\theta_{y}-\theta_{z} +\frac{1}{2}\right)\right]\widetilde{\Pi}=0, \nonumber\\ & &\left[\theta_{y} \left(\theta_{y} -\frac{1}{2}\right) - y\left(\theta_{y}+2\theta_{z}+\frac{1}{2}\right) \left(2\theta_{x}+\theta_{y} -\theta_{z}-\frac{1}{2}\right)\right]\widetilde{\Pi}=0, \nonumber\\ & &\left[\left(2\theta_{x}+\theta_{y}-\theta_{z}-\frac{1}{2}\right) \theta_{z} -z \left(\theta_{y} +2\theta_{z} +\frac{1}{2}\right) \left(\theta_{y}+2\theta_{z} +\frac{3}{2}\right) \right]\widetilde{\Pi}=0 \end{eqnarray} by $\Pi =x^{-1/8}z^{1/4}\widetilde{\Pi}$. An analytic solution around $(x,y,z)=(0,0,0)$ is given by \beq \widetilde{\Pi}=\sum_{m,n,p =0}^{\infty} \frac{(1/2)_{n+2p}(-1/2)_{2m+n-p}}{(1)_m (1/2)_n} \frac{x^m}{m!}\frac{y^n}{n!}\frac{z^p}{p!} ,\end{equation} which is known as the type $54b$ Srivastava and Karlsson's (Gaussian) hypergeometric function in three variables, \cite{SK} and we denote it by \beq G_{54b}[\alpha,\beta;\gamma,\delta;x,y,z]= \sum_{m,n,p =0}^{\infty} \frac{(\alpha)_{n+2p}(\beta)_{2m+n-p}}{(\gamma)_m (\delta)_n} \frac{x^m}{m!}\frac{y^n}{n!}\frac{z^p}{p!} ,\lab{223} \end{equation} which recovers Horn's ${\cal{H}}_4$ for $p=0$. Note that $G_{54b}$ satisfies \begin{eqnarray} & &\left[(\theta_{x} +\gamma -1)\theta_{x} -x\left(2\theta_{x}+\theta_{y}-\theta_{z}+\beta\right) \left(2\theta_{x}+\theta_{y}-\theta_{z} +\beta +1\right)\right]G_{54b}=0,\nonumber\\ & &\left[(\theta_{y} +\delta -1)\theta_{y} -y\left(\theta_{y}+2\theta_{z}+\alpha \right) \left(2\theta_{x}+\theta_{y} -\theta_{z} +\beta\right)\right]G_{54b}=0,\nonumber\\ & &\left[\left(2\theta_{x}+\theta_{y}-\theta_{z}+\beta\right) \theta_{z} -z \left(\theta_{y} +2\theta_{z} +\alpha \right) \left(\theta_{y}+2\theta_{z} +\alpha +1\right) \right]G_{54b}=0 .\end{eqnarray} \begin{center}
\section{Introduction} Since the discovery of superconductivity in alkali-metal doped $C_{60}$, extensive research on $C_{60}$ and other fullerenes has been carried out worldwide, aiming at understanding the mechanism for superconductivity and other related issues in fullerenes.\cite{varma,schluter,zhang,jishi,mitch,chakravarty,zhou,gunnarsson} Most of the theoretical models assumed that electron-phonon interaction is important for superconductivity.\cite{varma,schluter,zhang,jishi} Based on the analysis of the linewidths in vibronic spectra excited either by light (Raman scattering) or by neutrons, the electron-phonon coupling constant $\lambda$ for $A_3C_{60}$ (A=alkali metal) has been estimated. Recently, Winter and Kuzmany observed that the low frequency $H_g(1)$ and $H_g(2)$ modes lose all degeneracy and split into five components, each of which couples differently to the $t_{1u}$ electrons for single crystal of $K_3C_{60}$ at 80 K.\cite{winter} These results revealed that in the superconducting state, the pairing is mediated by phonons with weak or intermediate coupling. \cite{varma,schluter,mitch,zhou,gunnarsson,winter,guirion,prassides,koller} The lowest two unoccupied molecular orbitals of $C_{60}$ are both triply degenerated, having $t_{1u}$ and $t_{1g}$ symmetry. Filling of $t_{1u}$ and $t_{1g}$ bands with electrons is achieved by intercalation of alkali metals and alkaline earth metals to $C_{60}$ solids, respectively. Nevertheless, understanding of the "$t_{1g}$ superconductors" is extremely poor in comparison with the well known $t_{1u}$ superconductors. Comparison of physical property in between the $t_{1u}$ and $t_{1g}$ superconductors is of particular interest from the view point of mechanism of superconductivity. From the $t_{1u}$ symmetry of the electrons in the conduction band a coupling is only possible to the total symmetric $A_g$ modes and to the five-fold degenerate $H_g$ modes. While the coupling to the $A_g$ mode is expected to be weak due to an efficient screening effect, the $H_g$ modes may have a significantly strong coupling constant since they allow a Jahn-Teller mechanism. A similar coupling should take place in the case of the electrons with $t_{1g}$ symmetry. Superconductivity of Ba-doped $C_{60}$ was first discovered by Kortan et al, \cite{kortan} who claimed that the superconducting phase is bcc $Ba_6C_{60}$. Recently, Baenitz et al.,\cite{baenitz} on the other hand, reported that the superconducting phase is not $Ba_6C_{60}$ but $Ba_4C_{60}$. Very recently, we succeeded to synthesize single phase $Ba_4C_{60}$, and unambiguously confirmed that the $Ba_4C_{60}$ is the superconducting phase. In this work, we present results of a Raman scattering study of single phase $Ba_xC_{60}$ (x=3, 4 and 6 ) with $t_{1g}$ states. The results indicate that the electron-phonon interaction is also important for the $t_{1g}$ superconductor, particularly in superconducting $Ba_4C_{60}$. In addition, some amazing results were observed, particularly for the low frequency $H_g$ modes. (1) Raman shift of the tangential $A_g$ mode for $Ba_6C_{60}$ is much larger than the simple extrapolation relationship between Raman shift and charge transfer in alkali metal doped $C_{60}$; while the radial $A_g$ mode nearly remains unchanged with increasing charge transfer. (2) The Raman scattering behavior is quite different among the three phases of $Ba_3C_{60}$, $Ba_4C_{60}$ and $Ba_6C_{60}$, especially for the low frequency $H_g$ modes. The low frequency $H_g$ modes lose all degeneracy and split into five (or four) peaks at room temperature for the $Ba_4C_{60}$ and $Ba_6C_{60}$ samples, each of which couples differently to electrons with $t_{1g}$ symmetry. The splitting of low frequency $H_g$ modes into five components even at room temperature is similar to that observed in single crystal of $K_3C_{60}$ at low temperature of 80 K. \cite{winter} This is significant to understand the splitting and to evaluate the electron-phonon coupling constants for all directly coupling mode, estimating Tc in Ba-doped $C_{60}$. \section{Experiment} Samples of $Ba_xC_{60}$ (x=3, 4 and 6) were synthesized by reacting stoichiometric amount of powers of Ba and $C_{60}$. A quartz tube with mixed powder inside was sealed under high vacuum of about $2\times 10^{-6}$ torr. The samples of $Ba_3C_{60}$ and $Ba_6C_{60}$ were calcined at 600 $^oC$ for 216 hours with intermediate grindings of two times. In order to obtain high quality $Ba_4C_{60}$ sample, thermal annealing was carried out at 600 $^oC$ for 1080 hours with five intermediate grindings. X-ray diffraction showed that all samples were single phase, which is also confirmed by the single peak feature of the pentagonal pinch $A_g(2)$ mode in the Raman spectra. Raman scattering experiments were carried out using the 632.8 nm line of a He-Ne laser in the Brewster angle backscattering geometry. The scattering light was detected with a Dilor xy multichannel spectrometer using a spectral resolution of 3 $cm^{-1}$. Decomposition of the spectra into individual lines was made with a peak-fitting routine after a careful subtraction of the background originating from the laser. In order to obtain good Raman spectra, the samples were ground and pressed into pellets with pressure of about 20 $kg/cm^2$, which were sealed in Pyrex tubes under a high vacuum of $10^{-6}$ torr. \section{Results and Discussion} Figure 1 shows room temperature Raman spectra for the polycrystalline samples of $Ba_3C_{60}$, $Ba_4C_{60}$, and $Ba_6C_{60}$. For the three samples, only one peak of the pentagonal pinch $A_g(2)$ mode is observed, providing an evidence that each sample is in a single phase. These agree fairly well with the x-ray diffraction patterns. Interestingly, the three spectra have different strongest lines; they are $H_g(2)$, $A_g(1)$, and $A_g(2)$ modes for $Ba_3C_{60}$, $Ba_4C_{60}$, and $Ba_6C_{60}$, respectively. Another thing to be noted is that the half-width of all corresponding peaks of $Ba_4C_{60}$ is largest among $Ba_xC_{60}$ (x=3, 4 and 6 ) samples except for the $A_g(1)$ mode. This result is indicative of an importance of electron-phonon coupling in Raman spectrum of $Ba_4C_{60}$. Detailed discussion is given in the following. Also, it is to be pointed out that the Raman spectrum of $Ba_3C_{60}$ sample is amazingly similar to that of $K_6C_{60}$,\cite{zhou1} suggesting that the electronic states of $Ba_3C_{60}$ is similar to that of $K_6C_{60}$. This is in a fair agreement with a simple expectation that $C_{60}$ in both compounds is hexavalent. The frequency of the pentagonal pinch mode $ A_g(2)$ decreases with increasing Ba concentration, similarly to the case of alkali-metal doped $C_{60}$.\cite{duclos} The Raman shift of the $A_g(2)$ mode is discussed in the following. By contrast, the frequency of the radial $A_g(1)$ mode remains almost unchanged with Ba concentration, being different from the case of $K_xC_{60}$, where a slight up-shift of the radial $A_g(1)$ mode was observed.\cite{zhou1} The low frequency $H_g$ modes show dramatic changes depending on the Ba concentration. In particular, clear splittings are observed for the lowest frequency $H_g$ modes of $Ba_4C_{60}$ and $Ba_6C_{60}$. The positions ($\omega$) and halfwidths ($\gamma$) of the Raman modes observed are listed in Table I. For comparison, the lines for pure $C_{60}$ are included in Table I. In the following, we show detailed analysis of $H_g$ modes first, and then, discuss on the $A_g$ modes. In Fig.2 we show the results of a line-shape analysis of the Raman spectra of the $H_g(1)$ modes for $Ba_3C_{60}$, $Ba_4C_{60}$, and $Ba_6C_{60}$ samples. All modes were fit to a Lorentzian line shape. For $Ba_3C_{60}$ and $Ba_6C_{60}$, a doublet with Lorentzian components is observed, which has been observed in $K_6C_{60}$.\cite{zhou1} However, the $H_g(1)$ mode has to be fit with four components for $Ba_4C_{60}$. This splitting may be attributed to the symmetry lowering due to the orthorhombic structure of this material. A similar behavior has been observed in single crystal $K_3C_{60}$ at 80 K,\cite{winter} in which the $H_g(1)$ mode is split into five components. Position of the $H_g(1)$ components for $Ba_4C_{60}$ sample is nearly the same as that observed in $K_3C_{60}$. Figure 3 shows the higher resolution Raman spectra in the vicinity of 400 $cm^{-1}$ for $Ba_3C_{60}$, $Ba_4C_{60}$, and $Ba_6C_{60}$. While the cubic $Ba_3C_{60}$ shows a single peak at 432 $cm^{-1}$. $H_g(2)$ mode is apparently split into five components in $Ba_6C_{60}$. This splitting of $H_g(2)$ mode in $ Ba_6C_{60}$ is unexpected since the group theoretical consideration predicts a splitting into two in the space group $I_{m\overline{3}}$ ($T_5^h$). The splitting of the $H_g(2)$ mode might suggest a symmetry lowering which is not detected in the x-ray diffraction. This type of disagreement between microscopic spectroscopy and structural analysis was observed in $Rb_3C_{60}$, and still remains an open question. \cite{walstedt} A characteristic feature of the $H_g(2)$ mode of $Ba_6C_{60}$ is that the widths $\gamma$ of the components are almost the same except for the 428 $cm^{-1}$ component. By contrast, the $H_g(2)$ mode of $Ba_4C_{60}$ shows a strong peak at the high frequency edge associated with a long tailing structure towards lower frequencies. Linewidth and lineshift for the components are clearly related. A theoretical calculation shows the electron-phonon coupling constants are very sensitive to the change in the normal coordinates, the different components of the mode correspond to the different coupling constants.\cite{gunnarsson} It suggests that the fivefold degeneracy of the mode is lifted and each component couples with a different strength to the $t_{1g}$ carriers in $Ba_4C_{60}$. Results of a line-shape analysis of the Raman spectra of the $H_g(3)$ modes are shown in Fig.4. A doublet of $H_g(3)$ is observed for $Ba_3C_{60}$, which is ascribed to symmetry-lowering relative to $C_{60}$ molecules. \cite{zhou1} The $H_g(3)$ mode also displays a splitting into four both in $Ba_4C_{60}$ and $Ba_6C_{60}$. The splitting of the $H_g(3)$ mode in $Ba_6C_{60}$ also contradicts with the group theoretical consideration. It is to be pointed out that this anomalous splitting of $Ba_6C_{60}$ $H_g$ modes is observed only in $H_g(2)$ and $H_g(3)$ modes. The other $H_g$ modes are singlet or doublet, being consistent with the group theoretical consideration. In reference 9, Winter and Kuzmany gave several possible explanations for the splitting of the low frequency $H_g$ modes. (1) The splitting of the modes is understood from the merohedral disorder for the alkali derived metallic fullerides.\cite{fischer} This disorder is of low enough symmetry to allow only one dimensional representations for all modes. (2) The splitting originates from a Jahn-Teller type interaction. This interaction can give rise to a new vibrational system with rather large number of components, even more than five.\cite{auerbach} Also, a contribution to the splitting from an internal strain between the doped part of the crystal and the undoped part of the crystal. In our experiments, the low frequency $H_g$ modes almost lose all degeneracy for $Ba_4C_{60}$ and $Ba_6C_{60}$, and is different from that of $Ba_3C_{60}$ which is similar to that of $K_6C_{60}$ at room temperature. In the case of $Ba_4C_{60}$, the splitting can be understood since the crystal structure is orthorhombic. However, the splitting of $Ba_6C_{60}$ is not explained from the crystal structure. Particularly, when one considers that $Ba_6C_{60}$ is isostructural to $K_6C_{60}$, the splitting of $H_g(2)$ and $H_g(3)$ modes are considerably anomalous. This result might suggest that there exists a symmetry lowing which cannot be detected by x-ray diffraction. Similar symmetry lowing is observed in the NMR spectra of $Rb_3C_{60}$.\cite{walstedt} The next thing to be pointed out is that the splitting is observed even in polycrystalline samples and at room temperature, in contrast to the case of $K_3C_{60}$. In alkaline-earth-metal doped $C_{60}$, the local-density approximation calculations show a strong hybridization between the alkaline-earth-atom states and the $C_{60}$ $\pi $ states.\cite{saito,erwin} This hybridization, which is absent in the alkali-metal doped $C_{60}$, may play an essential role for the splitting of low frequency $H_g$ modes at room temperature. For the components of low frequency $H_g$ modes, a clear relation between line shift and line broadening is observed in $Ba_4C_{60}$, which is similar to that of single crystal $K_3C_{60}$. Winter and Kuzmany have pointed out that the electron-phonon interaction plays an important role in the broadening and the shift of the lines, and they deduced electron-phonon coupling constants.\cite{winter} The phonon linewidth broadening $\gamma_i$ due to the electron-phonon interaction in a metal can be related to a dimensionless electron phonon coupling constant $\lambda_i$ given by \cite{varma,allen} \begin{equation} \gamma_i = \frac{1}{g_i}\frac{\pi}{2}N(0)\lambda_i\omega_{bi}^2 \end{equation} where N(0) the density of states at the Fermi level per spin and molecule, and $g_i$ and $\omega_{bi}$ the mode degeneracy and the frequency before any coupling to the electrons, respectively. The Allen's formula given above will be used to derive the coupling constants for the eight $H_g$ modes. Frequencies of pure $C_{60}$ were used as the bare phonon frequencies. In the framework of Allen's theory there should be a linear relation of the form\cite{varma} \begin{equation} \gamma = -\frac{\pi }{2}N(0)\omega _b \Delta \omega \end{equation} between $\gamma$ the linewidth and $\Delta \omega$ the difference between the bare phonon frequency and the observed frequency. According to the experimental values of the three lowest frequency $H_g$ modes in Table I, the relations between linewidth and frequency shift is plotted in Fig.5 for $Ba_4C_{60}$ and $Ba_6C_{60}$ . The $\gamma$ and $\Delta \omega$ relation for $Ba_4C_{60}$ is linear and consistent with that expected from Eq.(2). N(0) can be deduced from the slope. The density of states obtained from the three $H_g$ modes are 7 eV$^{-1}$, 4 eV$^{-1}$ and 3.2 eV$^{-1}$, respectively. The discrepancy may arise from the fact that we could not use the real bare phonon frequencies for the evaluation. Geometry effects may also contribute to the shift and may be different for the modes. For $Ba_6C_{60}$, there exists no relation between the linewidth and lineshift in Fig.5b. N(0) is much less than 1 eV$^{-1}$ if it were deduced basing on the relation between linewidth and lineshift in Fig.5b. It suggests $Ba_6C_{60}$ could not follow electron-phonon coupling theory. It further supports that $Ba_4C_{60}$ is superconducting phase, rather than $Ba _6C_{60}$. For the evaluation of the coupling constants as discussed below a value of 7 $eV^{-1}$ is used for N(0). To our knowledge, no N(0) for $Ba_4C_{60}$ is available. The calculated N(0) is 4.3 states per eV\cite{saito} and an experimental value of 5.6 $eV^{-1}$ was reported for $Ba_6C_{60}$. \cite{gogia} The averaged linewidths and the overall coupling constants for each mode and for all $H_g$ modes for $Ba_4C_{60}$ are listed in Table II, together with the frequencies for the pure $C_{60}$. The averaged linewidths are directly evaluated from the linewiths listed in Table I. The values for $\lambda_i$ are evaluated using Eq.(1). The individual contributions to the coupling constant from each $H_g$ mode are listed in Table II. The three lowest frequency $H_g$ modes dominate the contribution to $\lambda$, yielding over 70\% of the total value. Large coupling constants of the low $H_g$ modes were also observed in $K_3C_{60}$.\cite{winter} Within the BCS framework, the superconducting transition temperature $T_c$ can been evaluated basing on the experimental values for $\lambda$ by the McMillan equation \begin{equation} T_{c}=\frac{\hbar\omega_{ln}}{1.2k_{B}}exp[\frac{-1.04(1+\lambda)} {\lambda-\mu^*-0.62\lambda \mu^*}] \end{equation} where $\omega_{ln}$ is the logarithmic averaged phonon frequency, $k_{B}$ is the Boltzmann constant, and $\mu^*$ is Coulomb repulsion between conduction electrons. According to the observed frequencies and the evaluated coupling constants, the $\omega_{ln}$ was determined as 490 $cm^{-1}$. With this value and $\lambda$, the superconducting transition temperature of 7 K can be evaluated, assuming the $\mu^*$ value as 0.3, however, which is anomalously large. The value for $\mu^*$ is much larger than 0.18 in $K_3C_{60}$ in the same way for evaluation of $T_c$. It might suggest a difference between $t_{1u}$ and $t_{1g}$ superconductors. To evaluate $T_c$, on the other hand, the logarithmic averaged phonon frequency of 150 $cm^{-1}$ is obtained if the $\mu^*$ is set as a reasonable value 0.2. In this case, the phonon frequency is significantly smaller than the intramolecular vibration range. Interestingly, the small phonon energy associated with superconductivity is also suggested by analysis of another $t_{1g}$ superconductor $A_3Ba_3C_{60}$.\cite{iwasa} Let us switch to the arguments on the totally symmetric $A_g$ modes. Figure 6 shows the Raman shift of the $A_g(2)$ pentagonal pinch mode as a function of nominal charge transfer simply derived from the chemical formula for $Ba_xC_{60}$. In this figure, we plotted the present results of $Ba_xC_{60}$, as well as that of $K_xC_{60}$ reported by Duclos et al. \cite{duclos} and the theoretical results of Jishi and Dresselhaus \cite{jishi1} for comparison. Since the plots of $Ba_xC_{60}$ approximately fall on an extrapolation of $K_xC_{60}$ or theoretical line, the charge transfer value from Ba to $C_{60}$ is almost complete. The molecular valences of $Ba_xC_{60}$ (x=3, 4, 6) are regarded as -6, -8, and -12, respectively. However, the situation is more complicated than the case of alkali doped materials. Several band calculations and experiments \cite{saito,erwin,niedrig} suggest a strong effect of hybridization of Ba and $C_{60}$ orbitals. If this is the case, the net charge transfer to $C_{60}$ is expected to be incomplete. In the present result, however, the charge transfer is approximately complete. Moreover, the slope of $Ba_xC_{60}$ is steeper than that of $K_xC_{60}$ or theory. These results indicate that the phonon mode should be reconsidered in the presence of metal-fullerene hybridization. Especially, there is a difference of 10 $cm^{-1}$ between the experimental and theoretical values for $Ba_6C_{60}$. The theory of Jishi and Dresselhaus focuses on the mode softening of the tangential vibrational $A_g$ mode for the alkali-metal derived fullerides, the hydridization between intercalants and $C_{60}$ was not considered. It can be seen from Table I that the frequency of the radial $A_g$ mode for $Ba_{3}C_{60}$ is 506 $cm^{-1}$. The upshift is as high as 13 $cm^{-1}$ relative to pure $C_{60}$. But, upon further doping with barium, the frequency nearly remains unchanged, being different from alkali-metal doped $C_{60}$, which shows a continuous hardening of $A_g(1)$ mode as a function of alkali metal concentration.\cite{zhou1} The mode-stiffening effect is due to electrostatic interactions which produces sufficient stiffening to encounter the softening of the mode expected on the basis of charge-transfer effects.\cite{jishi1,chen} In the case of Ba derived fullerides, there exists a strong hybridization between the Ba atoms and the $\pi$-type functions of the $C_{60}$ network. This may lead to a decrease in the electrostatic interactions, so that the frequency of the radial $A_{g}$ mode nearly remains unchanged with increasing Ba concentration. \section{Conclusion} Raman scattering studies of single phase $Ba_3C_{60}$, $Ba_4C_{60}$, and $Ba_6C_{60}$ have been carried out. The lowest frequency $H_g$ modes split in to five components for $Ba_4C_{60}$ and $Ba_6C_{60}$. A characteristic relation between lineshift and linewidth is observed in $Ba_4C_{60}$, this is consistent with that expected by electron-phonon interaction.While $Ba_6C_{60}$ does not exhibit such behavior. The characteristic relation is used to evaluate the N(0), the electron-phonon coupling constants are evaluated basing on the Raman results in the framework of Allen's theory. The radial $A_g$ mode shows a different behavior from alkali derived fullerides, the frequency remains unchanged with increasing Ba concentration; the effect of charge transfer on the softening of the tangential $A_g$ mode is larger in the alkaline-earth metal doped $C_{60}$ than in alkali derived $C_{60}$. These discrepancies may arise from the hybridization between intercalants and $C_{60}$ in alkaline-earth metal doped $C_{60}$. \section{acknowlegments} X. H. Chen would like to thank the Inoue Foundation for Science for financial support. This work is partly supported by Grant from the Japan Society for Promotion of Science (RFTF 96P00104, MPCR-363/96-03262) and from the Ministry of Education, Science, Sports, and Culture.
\section{Introduction} \label{sec:intro} Hysteresis often occurs when a bistable or multistable system is driven by an oscillating force which varies too fast for the system to respond without a phase lag. In fact, the word was coined by Ewing from the Greek {\it husterein\/} \mbox{($\stackrel{\scriptscriptstyle c}{\upsilon} \! \! \sigma \tau \epsilon \rho \acute{\epsilon} \omega$)} which means ``to be behind'' \cite{EWIN1881}. Examples include ferromagnets and ferroelectrics in AC fields, electrochemical cyclic-voltammetry experiments, and nonlinear elastic media under oscillating stress, just to mention a few. The most familiar representation of hysteresis is probably the (usually) closed curve obtained by plotting the system response versus the applied force. Examples of such {\it hysteresis loops\/} are shown in Fig.~\ref{fig:MHloops}(a). This figure shows data from a Monte Carlo (MC) simulation of a two-dimensional kinetic Ising ferromagnet below its critical temperature, which is driven by a sinusoidally oscillating field. However, the shape of the loops shown is quite general. For simplicity and concreteness, in this paper we use magnetic language, designating the oscillating force the ``field'' and the system response the ``magnetization''. These quantities can easily be re-interpreted when one discusses other types of systems, such as those mentioned above. Hysteresis was first systematically investigated in the late 19th century by engineers and physicists primarily concerned with the development of electric motors and transformers \cite{EWIN1881,WARB1881,EWIN1882,STEI1892,KLIN92}. The area of the hysteresis loop is proportional to the magnetic energy loss during one field period, as was first pointed out by Warburg \cite{WARB1881}. Its dependences on the frequency and amplitude of the applied field have therefore been studied intensively ever since, but open questions still remain. In particular, the question of whether the low-frequency behavior for ultrathin films of highly anisotropic materials is asymptotically a power law or logarithmic is still under investigation, both experimentally \cite{HE93,JIAN95,JIAN96,SUEN97,SUEN99} and theoretically \cite{JUNG90,RAO90B,THOM93,LUSE94,BEAL94,SIDE98B,SIDE99}. A different aspect of hysteresis in bistable systems, which is the main topic of the present paper, occurs at higher driving frequencies. When the frequency becomes sufficiently high, the symmetry of the hysteresis loop, which is apparent in Fig.~\ref{fig:MHloops}(a), is broken. Instead of oscillating between its two stable values with a phase lag relative to the field, the magnetization oscillates around one or the other of its zero-field stable values. A series of such asymmetric loops is shown in Fig.~\ref{fig:MHloops}(b). This symmetry breaking has become a topic of vigorous research during the last two decades. It was first reported by Tom{\'e} and de~Oliveira \cite{TOME90}, who observed it during numerical solutions of a mean-field equation of motion for a ferromagnet in an oscillating field. Subsequently, symmetry breaking has been observed in numerous MC simulations of kinetic Ising systems, \cite{SIDE99,LO90,ACHA95,ACHA97C,ACHA97D,ACHA98,SIDE98}, as well as in further mean-field studies \cite{ACHA95,ACHA97D,ACHA98,BUEN98}. It may also have been experimentally observed in ultrathin films of Co on Cu(001) \cite{JIAN95,JIAN96}. Reviews of the field as it stood in 1994 and as it stands today can be found in \cite{ACHA94} and~\cite{CHAK99}, respectively. There now appears to be a consensus that the symmetry breaking corresponds to a genuine second-order, nonequilibrium phase transition. Associated with the transition is a divergent time scale (critical slowing-down \cite{ACHA97D} and, for spatially extended systems, a divergent correlation length \cite{SIDE99,SIDE98}. Although estimates of the critical exponents have recently become available from finite-size scaling analyses of MC data for a two-dimensional Ising system in a sinusoidally oscillating field \cite{SIDE99,SIDE98}, their accuracy is not yet sufficient to decide whether the dynamic phase transition in this system belongs to a previously known universality class or represents a new one. Neither are we aware of theoretical arguments that can resolve the issue or help to identify the correct ``critical clusters''. The main purpose of this paper is to present preliminary results for two-dimensional kinetic Ising systems in square-wave oscillating fields. These may help point the way towards answering some of the remaining questions concerning this intriguing nonequilibrium critical phenomenon. The use of a square-wave field both tests the universality of the dynamic phase transition and significantly increases the computational speed. The structure of the remainder of the paper is as follows. The kinetic model system and the relevant quantities, including the dynamic order parameter, are defined in Sect.~\ref{sec:model}. A brief primer on metastable decay in spatially extended systems is given in Sect.~\ref{sec:KJMA}. Numerical results are presented in Sect.~\ref{sec:res}, and a discussion and some suggestions for further research are given in Sect.~\ref{sec:disc}. \section{Model and Relevant Quantities} \label{sec:model} Most numerical simulations of the dynamic phase transition (hereafter abbreviated DPT) in spatially extended systems have been performed on nearest-neighbor kinetic Ising ferromagnets on hypercubic lattices with periodic boundary conditions. These models are defined by the Hamiltonian \begin{equation} \label{eq:Hamil} {\cal H } = -J \sum_{ {\langle ij \rangle}} {s_{i}s_{j}} - H(t) \sum_{i} {s_{i}} \;, \end{equation} where $s_{i} \! = \! \pm 1$ is the state of the $i$th spin, $\sum_{ {\langle ij \rangle} }$ runs over all nearest-neighbor pairs, $J > 0$ is the ferromagnetic interaction, $\sum_{i}$ runs over all $L^{d}$ lattice sites, and $H(t)$ is an oscillating, spatially uniform applied field. The magnetization per site is \begin{equation} \label{eq:m(t)} {m(t)} = {L^{-d}} \sum_{i=1}^{L^d} {s_{i}(t)} \;. \end{equation} The temperature $T$ is fixed below its zero-field critical value $T_{\mathrm c}$, so that the magnetization for $H\!=\!0$ has two degenerate spontaneous equilibrium values, $\pm m_{\mathrm{sp}}(T)$. For nonzero fields the equilibrium magnetization has the same sign as $H$, while for not too strong $H$ (see Sect.~\ref{sec:KJMA} for quantitative statements) the opposite magnetization direction becomes {\it metastable\/} and decays slowly towards equilibrium with time. The dynamic used here, as well as in \cite{SIDE99} and \cite{SIDE98}, is the Glauber single-spin-flip MC algorithm with updates at randomly chosen sites \cite{BIND92B}. The time unit is one MC step per spin (MCSS). Each attempted spin flip from ${s_{i}}$ to ${-s_{i}}$ is accepted with probability \begin{equation} \label{eq:Glauber} W(s_{i} \rightarrow -s_{i}) = \frac{ \exp(- \beta \Delta E_{i})}{1 + \exp(- \beta \Delta E_{i})} \; . \end{equation} Here $\Delta E_{i}$ is the energy change that would result from accepting the spin flip, and $\beta \! = \! 1/k_{\rm B}T$ where $k_{\rm B}$ is Boltzmann's constant. For the largest system ($L \!=\! 512$ square lattice) we employed a massively parallel implementation of this algorithm \cite{LUBAA,LUBAB,KBNR,KNR}. Other dynamics that have been used in MC studies of the DPT are Glauber or Metropolis \cite{BIND92B} with updates at sequentially selected sites \cite{ACHA95,ACHA97C,ACHA97D,ACHA98,ACHA94,CHAK99}. Although the choice of update scheme can lead to subtle differences in the dynamics \cite{RIKV94A} and we prefer random site selection as the more physical scheme, we do not believe it affects universal aspects of the DPT. The dynamic order parameter is the period-averaged magnetization \cite{TOME90}, \begin{equation} \label{eq:Qeq} Q = \frac{1}{2 t_{1/2}} \oint m(t) \ {\mathrm d}t \;, \end{equation} where $t_{1/2}$ is the half-period of the oscillating field, and the beginning of the period is chosen at a time when $H(t)$ changes sign. Analogously we also define the local order parameter \begin{equation} \label{eq:Qlocaleq} Q_{i} = \frac{1}{2 t_{1/2}} \oint s_{i}(t) \ {\mathrm d}t \;, \end{equation} which is the period-averaged spin at site $i$. For slowly varying fields the probability distribution of $Q$ is sharply peaked at zero \cite{SIDE99}. We shall refer to this as the {\it dynamically disordered phase\/}. For rapidly oscillating fields the $Q$ distribution becomes bimodal with two sharp peaks near $\pm m_{\mathrm{sp}}(T)$, corresponding to the broken symmetry of the hysteresis loops \cite{SIDE99}. We shall refer to this as the {\it dynamically ordered phase\/}. Near the DPT we use finite-size scaling analysis of MC data to estimate the critical exponents that characterize the transition. Previous studies of the DPT have used an applied field which varies sinusoidally in time. While sinusoidal or linear saw-tooth fields are the most common in experiments and are necessary to obtain a vanishing loop area in the low-frequency limit \cite{SIDE98B,SIDE99}, the wave form of the field should not affect universal aspects of the DPT. This should be so because the transition essentially depends on the competition between two time scales: the half-period $t_{1/2}$ of the applied field, and the average time it takes the system to leave the metastable region near one of its two degenerate zero-field equilibria when a field of magnitude $H_0$ and sign opposite to the magnetization is applied. This {\it metastable lifetime\/}, $\langle \tau(H_0,T) \rangle$, is usually estimated as the average first-passage time to zero magnetization. In the present paper we use a {\it square-wave\/} field of amplitude $H_0$. This has significant computational advantages over the sinusoidal field variation since we can use two look-up tables to determine the acceptance probabilities: one for $H \!=\! + H_0$ and one for $H \!=\! - H_0$. In terms of the dimensionless half-period, \begin{equation} \label{eq:Theta} \Theta = t_{1/2} \left/ \langle \tau(H_0,T) \rangle \right. \;, \end{equation} the DPT should occur at a critical value $\Theta_{\mathrm c}$ of order unity. Although $\Theta$ can be changed by varying either $t_{1/2}$, $H_0$, or $T$, in a first approximation we expect $\Theta_{\mathrm c}$ to be independent of $H_0$ and $T$. This expectation is confirmed by simulations carried out at several $H_0$ and $T$ for different system sizes. In particular, Fig.~\ref{fig:Thetac}(a) shows the average norm of the order parameter $\langle |Q|\rangle$ vs.\ $\Theta$ for various field amplitudes and the corresponding metastable lifetimes. For weaker fields (longer lifetimes) the transition is apparent at some $\Theta_{\mathrm c} \sim 1$, while it clearly disappears (no dynamically ordered phase exists) for sufficiently strong fields (small lifetimes). Figure~\ref{fig:Thetac}(b) shows $\Theta_{\mathrm c}$ vs.\ $\langle \tau(H_0,T) \rangle$ where $\Theta_{\mathrm c}$ was determined approximately as the location of the peak in the fluctuations of $Q$ in a $L\!=\!64$ system. Note that $\Theta_{\mathrm c}$ for an infinite system is slightly different from this estimate due to finite-size effects. Also, these effects are expected to be smaller for strong fields (small lifetimes) where the spins become uncorrelated. The deviations from unity for very small $\langle \tau(H_0,T) \rangle$ and the complete disappearance of the DPT are discussed further in Sect.~\ref{sec:KJMA}. In many studies of the DPT the transition has been approached by changing $H_0$ or $T$ \cite{LO90,ACHA95,ACHA97C,ACHA98}. While the above discussion indicates that this is correct in principle, we show in Sect.~\ref{sec:KJMA} that $\langle \tau(H_0,T) \rangle$ depends strongly and nonlinearly on its arguments. We therefore prefer changing $t_{1/2}$ at constant $H_0$ and $T$ \cite{SIDE99,SIDE98}, as this in practice gives more precise control over the distance from the transition. \section{Decay of Metastable Phases} \label{sec:KJMA} In mean-field studies of the DPT $H_0$ must be larger than the temperature-dependent spinodal field beyond which the metastable free-energy minimum disappears \cite{JUNG90}. For weaker fields the hysteresis loops remain asymmetric, even for the lowest frequencies. It is much less appreciated that bounds on the fields and temperatures for which a DPT can occur also exist for spatially extended systems. These bounds are readily obtained from classical nucleation theory \cite{GUNT83B} and the Kolmogorov-Johnson-Mehl-Avrami (KJMA) theory of metastable decay \cite{SIDE99,SIDE98,RIKV94A,RIKV94,AVRAMI,RAMO99} by comparing four characteristic lengths. These are the lattice constant (here defined as unity), the size $R_{\rm c}(|H|,T)$ of a randomly nucleated critical droplet of stable phase, the typical distance $R_0(|H|,T)$ between individual droplets, and the system size $L$. The critical radius is determined by the competition between the surface free energy of the droplet, $\propto \sigma(T) R^{d-1}$ where $\sigma(T)$ is the surface tension between the two phases, and the bulk free energy, $\propto - |H| R^d$. As a result, $R_{\rm c} \propto \sigma(T)/|H|$. The nucleation rate per unit volume is $I(H,T) \propto \exp \left[ - \Xi(T) / |H|^{d-1} \right]$, where $\Xi(T)$ is the field-independent part of the free energy of a critical droplet divided by $k_{\rm B} T$, and preexponential powers of $|H|$ have been suppressed. The classical KJMA theory describes the metastable decay as homogeneous nucleation of droplets at random times and positions with nucleation rate $I$, followed by deterministic growth of these droplets with constant radial velocity $v \propto |H|$. Assuming that the droplets overlap freely when they meet, the time dependent magnetization is given by the well-known ``Avrami's Law'', \begin{eqnarray} m(t) &\approx& m_{\rm sp} \left\{ 2 \exp \left [- \int_{0}^{t} I \Omega_{d}(vt')^{d} {\rm d} t' \right ] - 1 \right\} \nonumber\\ &=& m_{\rm sp} \left\{ 2 \exp \left [- \frac{\Omega_{d} v^{d} I} {d+1} t^{d+1} \right ] - 1 \right\} \;, \label{eq:avrami} \end{eqnarray} where $\Omega_{d}$ is a proportionality constant such that the volume of a droplet of radius $R$ equals $\Omega_{d} R^{d}$. The argument of the exponential is the ``extended volume'' \cite{AVRAMI}, i.e., the total volume fraction of stable-phase droplets, {\it uncorrected\/} for overlaps. Solving (\ref{eq:avrami}) for the time at which $m \! = \! 0$ gives the lifetime, \begin{equation} \langle \tau (|H|,T) \rangle = \left [ \frac{\Omega_{d} v^{d} I} {(d+1) \ln 2 } \right ]^{-{1}/({d+1})} , \label{eq:tau_md} \end{equation} which depends on $|H|$ and $T$ through $v$ and $I$, but is {\it independent\/} of $L$. The characteristic length $R_0$ is obtained from $v(|H|,T)$ and $\langle \tau (|H|,T) \rangle$ as \begin{equation} R_0(|H|,T) = v(|H|,T) \ \langle \tau (|H|,T) \rangle \propto \exp \left[\frac{\Xi(T)}{(d+1) |H|^{d+1}} \right] \;, \label{eq:R0} \end{equation} where a preexponential power of $|H|$ again has been suppressed \cite{RIKV94A,RIKV94}. The regime in which a DPT can occur is that in which a large number of droplets contribute to the decay of the metastable phase: \begin{equation} 1 \ll R_{\rm c} \ll R_0 \ll L \;. \label{eq:lengths} \end{equation} This is known as the multidroplet (MD) regime \cite{RIKV94A,RIKV94}. It is limited on the weak-field/small-system side by the {\it dynamic spinodal\/} (DSP) field , \begin{equation} H_{\rm DSP}(T,L) \sim \left( \frac{1}{d+1} \frac{\Xi(T)}{\ln L}\right)^{{1}/({d-1})} \;, \label{eq:HDSP} \end{equation} which corresponds to $R_0 \! \approx \! L$. For $|H| < H_{\rm DSP}$ almost always only a {\it single\/} droplet contributes to the magnetization switching. A DPT does {\it not\/} occur in this regime, but {\it stochastic resonance\/} is observed at low frequencies \cite{SIDE98A}. On the strong-field side the MD regime is limited by an $L$-independent cross-over field approximately given by $2 R_{\rm c} \! \approx \! 1$. By a somewhat confusing term this crossover is often referred to as the ``mean-field spinodal'' $H_{\rm MFSP}(T)$ \cite{RIKV94A,RIKV94}. In the {\it strong-field\/} (SF) regime beyond this limit the spins become increasingly uncorrelated as $|H|$ increases. This is the region of small $\langle \tau(H_0,T) \rangle$ (large $H_0$) where $\Theta_{\rm c}$ rapidly approaches zero and the transition disappears (Fig.~\ref{fig:Thetac}). In computational studies of the DPT it is essential to ensure that $H_{\rm DSP}(T,L) < H_0 < H_{\rm MFSP}(T)$ for all values of $L$ and $T$ used. \section{Results} \label{sec:res} We performed extensive simulations on square lattices with $L$ between 64 and 512 at $T \!=\! 0.8T_{\rm c}$ and $H_0 \!=\! 0.3J$. Typical run lengths near the transition range from $2$$\times$$10^4$ to $10^5$ full periods, corresponding to $2.8$$\times$$10^6-1.4$$\times$$10^7$ MCSS. The system was initialized with all spins up and and the square-wave external field started with the half-period in which $H\!=\!-H_0$. After some relaxation the system magnetization reaches a limit cycle (except for thermal fluctuations), i.e., $Q$ is stationary. We discarded the first 500 periods of the time series to exclude transients from the stationary-state averages. For small half-periods ($\Theta\ll\Theta_{\rm c}$) the magnetization does not have time to switch, resulting in $|Q| \! \approx \! m_{\rm sp}$, while for large half-periods ($\Theta\gg\Theta_{\rm c}$) it switches every half-period and $Q\! \approx \! 0$ as can be seen from the time series in Fig.~\ref{fig:Qseries}. The transition between the high- and low-frequency regimes is singular, characterized by large fluctuations in $Q$. To illustrate the spatial aspects of the transition we also show configurations of the local order parameter $Q_i$ in Fig.~\ref{fig:configs}. Below $\Theta_{\rm c}$ [Fig.~\ref{fig:configs}(a)] the majority of spins spend most of their time in the $+1$ state, i.e., in the metastable phase during the first half-period, and in the stable equilibrium phase during the second half-period (except for equilibrium fluctuations). Thus, most of the $Q_i \! \approx \! +1$. Droplets of $s_i \!=\! -1$ that nucleate during the negative half-period and then decay back to $+1$ during the positive half-period show up as roughly circular gray spots in the figure. Since the spins near the center of such a droplet become negative first and revert to positive last, these spots appear darkest in the middle. Above $\Theta_{\rm c}$ [Figs.~\ref{fig:configs}(c,d)] the system follows the field in every half-period (with some phase lag) and $Q_i \! \approx \! 0$ at all sites $i$. Near $\Theta_{\rm c}$ [Fig.~\ref{fig:configs}(b)] there are large clusters of both $Q_i \! \approx \! +1$ and $Q_i \! \approx \! -1$ separated by wide ``interfaces'' where $Q_i \! \approx \! 0$. Also, for not too large lattices one often observes the full reversal of an ordered configuration $\{Q_i\}\rightarrow -\{Q_i\}$, typical of finite, spatially extended systems undergoing symmetry breaking. For finite systems in the dynamically ordered phase the distribution of $Q$ becomes bimodal. Thus, to capture symmetry breaking, one has to measure the average norm of Q as the order parameter, i.e., $\langle |Q| \rangle$. Figure~\ref{fig:fss}(a) clearly shows that this order parameter is of order unity for $\Theta<\Theta_{\rm c}$ and vanishes for $\Theta>\Theta_{\rm c}$, except for finite-size effects. To characterize and quantify this transition in terms of critical exponents we employ the well-known technique of finite-size scaling \cite{BIND92B,BIND90}. The quantity analogous to the susceptibility is the scaled variance of the dynamic order parameter, \begin{equation} X_L=L^2 \left(\langle Q^2 \rangle_L - \langle |Q| \rangle_L ^2 \right) \; . \label{eq:X} \end{equation} For finite systems $X_L$ has a characteristic peak near $\Theta_{\rm c}$ [see Fig.~\ref{fig:fss}(b)] which increases in height with increasing $L$, while no finite-size effects can be observed for $\Theta\ll\Theta_{\rm c}$ and $\Theta\gg\Theta_{\rm c}$. This implies the existence of a divergent length scale, possibly the correlation length which governs the long-distance behavior of the local order-parameter correlations $\langle Q_i Q_j \rangle$. Note that the location of the maximum in $X_L$ shifts with $L$. This also contains important information about the critical exponents. To estimate the value of $\Theta_{\rm c}$ at which the transition occurs in an infinite system we use the intersection of the fourth-order cumulant ratios \cite{BIND92B,BIND90}, \begin{equation} U_L=1 - \frac{\langle Q^4\rangle_L}{3\langle Q^2\rangle_L^2} \;, \label{eq:cumulant} \end{equation} for several system sizes as shown in Fig.~\ref{fig:fss}(c). For the largest system ($L \!=\! 512$) the error bars on $U_L$ were too large to use it to obtain estimates for the crossing. Our estimate for the dimensionless critical half-period is $\Theta_{\rm c} \!=\! 0.913 \pm 0.003$ with a fixed-point value $U^* \!=\! 0.615 \pm 0.005$ for the cumulant ratio. For our model the quantity analogous to the reduced temperature in equilibrium systems (i.e., the distance from the critical point) is \begin{equation} \theta= \frac{|\Theta-\Theta_{\rm c}|}{\Theta_{\rm c}}\;. \label{eq:reduced_Theta} \end{equation} Finite-size scaling theory provides simple relations for the order parameter and its scaled variance $X_L$ for finite systems in the critical regime \cite{BIND92B,BIND90}: \begin{subeqnarray} \langle |Q|\rangle _L & = & L^{-\beta /\nu} {\cal F}_{\pm}(\theta L^{1/\nu}) \;, \label{eq:Qscaling1} \\ X_L & = & L^{\gamma /\nu} {\cal G}_{\pm}(\theta L^{1/\nu})\;, \label{eq:Xscaling1} \end{subeqnarray} where ${\cal F}_{\pm}$ and ${\cal G}_{\pm}$ are scaling functions and the $+$ ($-$) index refers to \mbox{$\Theta >\Theta_{\rm c}$} \mbox{($\Theta <\Theta_{\rm c}$)}. Then at $\Theta_{\rm c}$ ($\theta\!=\!0$) we have the finite-size behavior of the above two observables: \begin{subeqnarray} \langle |Q|\rangle _L & \propto & L^{-\beta /\nu} \;, \label{eq:Qscaling2} \\ X_L & \propto & L^{\gamma /\nu} \;. \label{eq:Xscaling2} \end{subeqnarray} Using $\langle |Q|\rangle _L$ and $X_L$ for several system sizes at $\Theta_{\rm c}$ (estimated as the value of $\Theta$ where the cumulants cross) we employ (\ref{eq:Qscaling2}) and (\ref{eq:Xscaling2}) to find the exponent ratios $\beta /\nu$ and $\gamma /\nu$ through weighted linear least-squares fitting to the logarithmic data. Further, the shift in the location of the peak in $X_L$ for these finite systems, $\Theta_{\rm c}(L)$, is used to estimate $\nu$ in the same way \cite{BIND92B,BIND90}: \begin{equation} |\Theta_{\rm c}(L)-\Theta_{\rm c}| \propto L^{-1/\nu} \;. \label{eq:Theta_cscaling} \end{equation} Our estimates for these exponents are $\beta/\nu$$=$$0.122\pm 0.005$, $\gamma/\nu$$=$$1.77\pm 0.05$, and $\nu$$=$$1.0\pm 0.15$. The largest uncertainty occurs in $\nu$ due to the large relative errors in $|\Theta_{\rm c}(L)-\Theta_{\rm c}|$. We note that these numbers are very close (within the one-standard-deviation error bars) to the critical exponents of the two-dimensional Ising universality class $\beta/\nu$$=$$1/8$$=$$0.125$, $\gamma/\nu$$=$$7/4$$=$$1.75$, and $\nu$$=$$1$. However, as discussed in Sect.~\ref{sec:disc}, we do not consider this conclusive proof of the universality class for this nonequilibrium phase transition. To obtain a more general picture of how well the scaling relations in (\ref{eq:Qscaling1}) and (\ref{eq:Xscaling1}) are obeyed, we plot $\langle |Q|\rangle _L L^{\beta/\nu}$ [Fig.~\ref{fig:ising_collapse}(a)] and $X_L L^{-\gamma/\nu}$ [Fig.~\ref{fig:ising_collapse}(b)] vs. $\theta L^{1/\nu}$ \cite{LANDAU76}. In these figures we used the exponents of the two-dimensional Ising universality class since they are within one standard deviation of our estimates and analogous data-collapse plots using our numerical exponent estimates do not look perceptibly different. The plots show excellent agreement with the scaling assumption and graphically define the scaling functions ${\cal F}_{\pm}$ and ${\cal G}_{\pm}$. The asymptotic behaviors of these functions have to be simple power laws so that the true critical behavior is restored when the limit $L\rightarrow\infty$ is taken in (\ref{eq:Qscaling1}) and (\ref{eq:Xscaling1}). Some deviation can be observed in Fig.~\ref{fig:ising_collapse}(b): below $\Theta_{\rm c}$ the data points for the smaller systems systematically start to peel off earlier from the straight line representing the asymptotic behavior of ${\cal G}_{-}$ for large argument. In the absence of theoretical arguments why this DPT should belong to the Ising universality class we also tested the data collapse with the exponents for random percolation, which are relatively close to the corresponding Ising ones \cite{STAU92}: $\beta/\nu$$=$$5/48$$\approx$$0.104$, $\gamma/\nu$$=$$43/24$$\approx$$1.79$, and $\nu$$=$$4/3$$\approx$$1.33$. It appears reasonable to consider this universality class in particular, since the metastable decay process described in Sect.~\ref{sec:KJMA} produces transient spanning clusters that belong to the random-percolation class as $m(t)$ passes through $H$- and $T$-dependent percolation thresholds for the two phases near $m \! = \! 0$ \cite{SIDEUNP}. Data-collapse plots using the random-percolation exponents are shown in Figs.~\ref{fig:ising_collapse}(c,d). They are clearly inferior to the ones with the Ising exponents [Figs.~\ref{fig:ising_collapse}(a,b)]. In particular, for $\Theta >\Theta_c$ (the dynamically disordered phase) the data collapse gets progressively worse for larger systems. Comparing the scaling plots in Figs.~\ref{fig:ising_collapse}(a,b) with Figs.~\ref{fig:ising_collapse}(c,d) we conclude that the exponent $\nu$ for the DPT is significantly different from that of the random percolation universality class and that it is closer to the Ising value. This conclusion is supported by direct comparison of the random-percolation and Ising exponents with our numerical estimates for the DPT. Our estimates for $\beta/\nu$ and $\nu$ lie more than two standard deviations away from the percolation values, but less than one standard deviation away from the Ising values, while our estimate for $\gamma/\nu$ lies midway between and less than one standard deviation away from the exact values for both classes. The exponents previously obtained with a sinusoidally oscillating field \cite{SIDE99,SIDE98A} lie within two standard deviations of our values. As it was difficult to determine the uncertainties in the estimates obtained in that study, we believe the results are consistent and indicate that the DPT in systems driven by sinusoidal and square-wave fields belong to the same universality class, as expected. \section{Summary and Discussion} \label{sec:disc} In this paper we have presented preliminary results from a large-scale Monte Carlo study of the dynamic phase transition (DPT) in a two-dimensional kinetic Ising ferromagnet driven by a square-wave oscillating field. Our results are consistent with those of previous studies of the same model in a sinusoidally varying field \cite{SIDE99,SIDE98}. They indicate that for field amplitudes $H_0$ such that the metastable magnetized phase decays to equilibrium via the multidroplet (MD) mechanism described in Sect.~\ref{sec:KJMA}, the system undergoes a continuous DPT when the half-period of the field, $t_{1/2}$, approximately equals the metastable lifetime, $\langle \tau(H_0,T) \rangle$. Thus the critical value $\Theta_{\rm c}$ of the dimensionless half-period defined in (\ref{eq:Theta}) is near unity. As $\Theta$ is increased past $\Theta_{\rm c}$ the order parameter $\langle |Q| \rangle$, which is the expectation value of the norm of the period-averaged magnetization, vanishes in a singular fashion, as shown in Fig.~\ref{fig:fss}(a). The strong and systematic finite-size effects in the order parameter and its scaled variance indicate that there is a divergent correlation length associated with the transition. Using standard finite-size scaling techniques borrowed from the theory of equilibrium phase transitions we therefore estimated $\Theta_{\rm c}$ and the critical exponents $\beta$, $\gamma$, and $\nu$ from data for system sizes between $L \!=\! 64$ and 512. The resulting estimates are $\beta/\nu$$=$$0.122\pm 0.005$, $\gamma/\nu$$=$$1.77\pm 0.05$, and $\nu$$=$$1.0\pm 0.15$, which agree to within the statistical errors with those previously obtained with a sinusoidally oscillating field \cite{SIDE99,SIDE98A}. This is strong evidence that the shape of the wave does not effect the universal aspects of the DPT. We also note that our estimates agree well within one standard deviation with the exact values for the two-dimensional equilibrium Ising model: $\beta/\nu$$=$$1/8$$=$$0.125$, $\gamma/\nu$$=$$7/4$$=$$ 1.75$, and $\nu$$=$$1$, and that they satisfy the exponent relation \begin{equation} 2 \left( \beta / \nu \right) + \left( \gamma / \nu \right) = 2.01 \pm 0.05 \approx d \;. \label{eq:hyper} \end{equation} The fixed-point value of the fourth-order cumulant ratio, $U^* \!=\! 0.615 \pm 0.005$, is also close to that of the Ising model, $U^* \!=\! 0.610 \ 690 \ 1(5)$ \cite{KAMI93}. As discussed in \cite{SIDE99} we find the estimated values for the exponents and $U^*$ conclusive evidence that the DPT corresponds to a nontrivial fixed point, so that (\ref{eq:hyper}) indeed represents a hyperscaling relation. Since we know of no convincing theoretical argument that this nonequilibrium phase transition should be in any particular equilibrium universality class, we also compare the exponent estimates with those of the universality class of random percolation. For this class the values \cite{STAU92} $\beta/\nu$$=$$5/48$$\approx$$0.104$ and $\nu$$=$$4/3$$\approx$$1.33$ are significantly different from our exponent estimates, while $\gamma/\nu$$=$$43/24$$\approx$$1.79$ lies within the uncertainty in our estimate. Taken together, the three estimates and the low quality of the data collapse plots shown in Figs.~\ref{fig:ising_collapse}(c,d) constitute strong evidence that the DPT is {\it not} in the universality class of random percolation. However, we emphasize that we do not yet consider the question of the universality class as settled. That will require still more accurate numerical exponent estimates as well as theoretical arguments. Outside the MD regime the phase transition disappears. For weaker fields or smaller systems the metastable phase decays stochastically through a single droplet, and no long-range correlations evolve \cite{SIDE98A}. {}In the other extreme, as $H_0$ is increased into the strong-field regime, $\Theta_{\rm c}$ goes to zero and vanishes at a sharply defined field (lifetime), as shown in Fig.~\ref{fig:Thetac}(b). In conclusion, the dynamic phase transition observed in spatially extended kinetic Ising systems driven by oscillating fields is a fascinating nonequilibrium critical phenomenon with all the hallmarks of a thermal phase transition corresponding to a nontrivial fixed point. Although the present study confirms that the particular wave form of the oscillating field does not change the universality class, and our numerical estimates for the critical exponents and cumulant ratio are consistent with the universality class of the two-dimensional Ising model in equilibrium, many questions remain. These include the question of universality under changes in $T$ and $H_0$, and the details of the disappearance of the transition as $H_0$ is increased. No procedure to identify critical correlated clusters from the connected clusters seen in configuration ``snapshots'' such as Fig.~\ref{fig:configs}(b) is yet known, and theoretical ideas are largely missing. Thus, there is still much to be done! \section*{Acknowledgments} \label{sec:ack} This research was supported in part by Florida State University through the Center for Materials Research and Technology and the Supercomputer Computations Research Institute (partially funded by the US Department of Energy through Contract No.\ DE-FC05-85ER25000), and by the National Science Foundation through Grants No.\ DMR-9634873 and DMR-9871455. Computing resources at the National Energy Research Scientific Computing Center were made available by the US Department of Energy under Contract No.\ DE-AC03-76SF00098.
\section{Introduction} It is no exaggeration to say that between redshifts of $\approx$ 2-3 to the present, highly luminous (M $<$ $-$26) quasars have become virtually extinct. The co-moving space density of luminous quasars has fallen by about a factor of 1000 between these two epochs ({\it e.g.,\ } Hartwick \& Schade 1990; Boyle 1993). What processes led to such a strong density evolution of highly luminous active galactic nuclei (AGN)? The answer to this question not only has important ramifications for our understanding of the AGN phenomenon, but will almost certainly give us important insight into the processes that controlled the evolution of galaxies in general since the \lq\lq quasar epoch\rq\rq . It is an intriguing possibility that there is a strong link between the fueling of luminous AGN and galaxy formation. Indeed, the quasar epoch occurs at the time \lq\lq Cold Dark Matter\rq\rq \ models identify as when typical present-day galaxies where hierarchically assembled via dissipative mergers. Rees (1988), without regard to any specific cosmogony, argues that the process of galaxy formation was still continuing at the epoch z=2. This is especially true of the theoretical models where galaxy formation is a slow, diffuse, low-luminosity process (e.g., Baron \& White 1987; Kauffmann, White, \& Guiderdoni 1993; Baugh et al.~1998). These general speculations therefore give a special prominence to high redshift AGN by hypothesizing that only when the proto-galactic material is energized by the luminosity of an AGN will it have high enough surface brightness to be readily detectable in emission-line surveys. In turn, the process of galaxy formation may be substantially altered by the effect of the AGN. Observations of radio-loud quasars are also valuable for what we can learn about the AGN phenomenon in general. In particular, the relationship between radio-loud quasars and radio galaxies is of considerable interest, especially in light of efforts to \lq\lq unify\rq\rq \ these two classes through differences in viewing angle, environment, or evolutionary state ({\it e.g.,\ } Barthel 1989; Norman \& Miley 1984; Neff \& Hutchings 1990). In particular, the \lq\lq viewing angle\rq\rq \ scheme of Barthel (1989) in which radio-loud quasars and radio galaxies are drawn from the same parent population but viewed preferentially at small or large angles, respectively, to the radio axis, predicts that the luminosity and color of the quasar host should be very similar, if not identical, to those of the radio galaxies at similar radio powers and redshifts. Moreover, radio galaxies at high redshifts (z $>$ 0.6) exhibit the so-called \lq\lq alignment effect\rq\rq \ (McCarthy et al.~1987; Chambers et al.~1987) where the radio, and the rest-frame UV and optical axes are all roughly co-linear. Recently, through the use of broad-band HST imaging data, it has become clear that quasars do indeed exhibit the ``alignment effect'' (Lehnert et al.~1999a) and that the effect in quasars is weaker than that seen in radio galaxies, as expected in simple orientation-based unification schemes (Lehnert et al.~1999b in preparation). Radio-loud quasars at redshifts of 2-3 have impressively large and luminous continuum and emission-line nebulae (``hosts''; Heckman et al.~1991a,b; Lehnert et al.~1992). These structures comprise about 3\% to over 40\% of the total flux from the quasar and have Ly$\alpha$ luminosities of $\approx$10$^{44-45}$ ergs s$^{-1}$, absolute visual magnitudes of about $-$25, blue spectral energy distributions consistent with those of nearby late-type galaxies (Sc and Irr), and sizes of the order of many tens of kiloparsecs. Ground-based images of high-z radio loud quasars contain only limited morphological information (1 arc second seeing corresponds to about 11 kpc at these redshifts) and their morphology can best be described as asymmetric. Host galaxies of low-redshift quasars ({\it e.g.,\ } Boroson \& Oke 1984; Boroson, Persson, \& Oke 1985; Smith et al.~1986) have emission-line luminosities much lower than this (about 10$^{42}$ ergs s$^{-1}$ in [OII]), spectral energy distributions similar to late-type galaxies, and absolute magnitudes of around $-$22. However, Bahcall et al.~(1994, 1995a,b) have recently concluded using HST WFPC2 images that the host galaxies of some of the brightest, low-redshift QSOs have wide range of host luminosities, with perhaps a large fraction of quasars hosts having luminosities $< L_{\star}$. However, subsequent re-analysis of the data show that the host galaxies of low redshift quasars are in fact large luminous galaxies (McLeod \& Rieke 1995; Bahcall et al.~1997). In spite of this, the properties of host galaxies of quasars have clearly evolved very strongly from high-redshifts (Kotilainen, Falomo, \& Spencer 1998; Lehnert et al.~ 1992). As tantalizing as the data and information contained in these studies of high redshift quasars are, there are some serious limitations to the usefulness of our previously obtained ground-based data. It is essential to be able to separate the strong nuclear emission of the quasar (continuum and Ly$\alpha$) from that of its surrounding host. This can be most easily accomplished with the HST. Accurate separation of the quasar nucleus from its surrounding emission (within 10-20 kpc) allows us to address a number of issues which are of importance to our understanding of the evolution of luminous quasars, the association and similarity between high-redshift radio galaxies and radio-loud quasars, and the role of the large-scale radio emission in exciting and enhancing the extended line emission. To these ends, we used the HST to study a small subset of high-redshift, radio-loud quasars from our ground-based sample (Heckman et al.~1991a). In this paper we represent an analysis of these observations, using several different methods to test the robustness of our results, and compare the quasar host properties with those of radio galaxies. Throughout this paper we adopt H$_0 = 50$ km s$^{-1}$ Mpc$^{-1}$, q$_0 = 0.1$, and $\Lambda = 0$ for easy comparison to previously published work on high redshift radio galaxies. \section{Observations and Reduction} The HST observations of our sample were made from September 1994 through May 1995 using the Wide Field Planetary Camera 2 (WFPC2). Each quasar was centered on the PC and was observed for a total of 2100 seconds (3 $\times$ 700 seconds) through the F555W filter (which closely corresponds to a ground-based V filter). In addition, we obtained exposures of 5000 seconds (5 x 1000 seconds) using the Wide Field Camera (WFC) and one of the Quad [OII] filters. The narrow-band filter was selected such that its central wavelength corresponded closely to the wavelength of redshifted Ly$\alpha$ from each of the quasars. The observations are summarized in Table 1. The pixel size is 0.046\arcsec \ pixel$^{-1}$ in the PC and 0.1\arcsec \ pixel$^{-1}$ in the WFC. The individual exposures were reduced using the standard pipeline reduction. The final images were produced by averaging the individual exposures with sigma-clipping to remove cosmic rays. The bandpass of the telescope + WFPC2 + F555W combination covers the wavelength range of approximately 4500\AA \ and 6000\AA . The strong UV lines of CIV$\lambda$1550, HeII$\lambda$1640, and CIII]$\lambda$1909 are included in this bandpass and in principle it is possible that these lines contribute to the detected extended emission. However, Heckman et al.~(1991$b$) and Lehnert \& Becker (1998) have shown that these lines are generally weak in quasar hosts, with the possible exception of PKS 0445+097. For this quasar Heckman et al.~(1991$b$) found that the host has a HeII$\lambda$1640 equivalent width of several \AA. From this and the brightness of the host in our PC image we estimate that HeII emission contributes much less than 1\% to the total emission seen in the PC image. Thus the contribution of line emission to the total emission observed from the host in the F555W filter is negligible, for PKS 0445+097 and all other quasars in our sample. The broad-band data were flux-calibrated assuming the inverse sensitivity for the F555W filter of 3.459$\times$ 10$^{-18}$ ergs s$^{-1}$ cm$^{-2}$ \AA$^{-1}$ dn$^{-1}$ and a zero point of 22.563 (Whitmore 1995). This puts the resultant magnitudes on the ``Vega system''. To convert to the STMAG system, which assumes a flat spectral energy distribution and has a zero point of 22.543, 0.020 magnitudes should be subtracted from the magnitudes given here. \section{Image Reduction} \subsection{Construction and Systematics of the PSF} We have attempted to quantify the shape and constancy of the HST PSF. First, we have collected images of the standard stars used to calibrate the F555W filter and observations of bright stars in the outer regions of $\omega$ Cen that were made within a period of days of our observations. We found about 20 stars that were suitably close in time and within about 100 pixels of the central pixel of the PC (all of quasar images where centered near the middle of the PC CCD). We then constructed an empirical point-spread function using these data by adding the individual exposures after they had been aligned to a common center. This empirical PSF was then compared with the model PSF constructed using the PSF modeling program, ``Tiny Tim''. However, the close agreement is limited to azimuthal averages -- Tiny Tim does not reproduce the detailed 2-dimensional structure of the PSF (there is an asymmetry in the intensity of the diffraction spikes, especially in the $+$U3 direction, which Tiny Tim does not reproduce well). Next, we also measured the encircled energy diagrams (EEDs), i.e., the fraction of flux from a point source interior to a radius r, as a function of r. We then inter-compared all the EEDs taken through a given filter to determine the reproducibility of the EED and compared the individual stellar EEDs with that of the model PSF produced by Tiny Tim. We found very good agreement between the shape of the EED from the sum of the observations of standard stars and that of the Tiny Tim PSF. We then compared individual star EEDs with that of the Tiny Tim PSF. This inter-comparison of approximately 20 stars showed that we can detect host that contributes more than about 5\% as much light as the quasar itself (within a radius of about 1.4\arcsec). This limit is consistent with the known temporal variations in the HST PSF due to effects like the gentle change in focus over times scales of months and shorter time scale variations due to the so-called ``breathing'' of the telescope (see Burrows et al.~1995). We have restricted ourselves to radii less than about 1.5\arcsec , to avoid the effect of the poorly understood large angle scattering which becomes important beyond a radius of about 2\arcsec. \subsection{Host Measurements: PSF Subtraction} Measuring the properties of the circum-nuclear emission is important if we are to gain true insight into the properties of quasar host galaxies. One arc second at the redshift of quasars corresponds to roughly 11 kpc (in the cosmology: H$_0$=50 km s$^{-1}$ Mpc$^{-1}$ and q$_0$=0.1 which will be used throughout this paper) -- which is similar to the sizes of present-day galaxies. Thus being able to accurately subtract the nuclear emission of the quasar from the more extended emission is critical if we are to obtain information about the host of high redshift quasar on scales a fraction of the present-day galaxy size. We attempted this subtraction using two different PSFs. In the first, we modeled the PSF using the HST PSF modeling program, Tiny Tim. The second method was to construct a PSF by averaging several images (40s integrations) of a bright star in the outer regions of $\omega$ Centauri using the F555W and the PC. These stellar images were all within about 100 pixels of the center of the PC chip and thus near the location of the quasar image on the PC chip. The images were aligned to the nearest pixel before averaging and the individual images covered the central 200 pixels of the planetary camera. The counts in the central pixel was about 2500 DN, comparable to that in the images of the quasars. We restricted ourselves to these observations and not the entire 20 that were used in the EED analysis. The subsample we used were chosen to sample the range of PC positions covered by the F555W quasar exposures as part of this program. These two PSFs were iteratively subtracted until emission due to the diffraction of the secondary support was zero. This procedure allowed us to estimate the uncertainty in our fraction of extended emission by observing the points were the diffraction spikes became negative due to over subtraction of the PSF or were obvious in the PSF subtracted image due to under subtraction. We estimate that the range of possible acceptable amounts of PSF subtraction lead to a factor of 30\% or $\pm$0.3 magnitudes in the quoted flux from the host. We note that these diffraction spikes were seen only out to about 1\arcsec \ in the PC frames and there was very little underlying nuclear emission from the quasar over most of this area. Using the model from Tiny Tim or the images of the star in $\omega$ Cen gave very similar results ({\it i.e.,\ } very similar fraction of the emission that was extended and similar morphology of the underlying emission). Therefore throughout the remainder of the paper, we will quote only the results of PSF subtraction obtained by subtracting the image of the star through the F555W and the PC. The narrow-band images of the quasars generally only reveal modest extended emission. Using the efficiencies of the telescope, detector and filter combination made available by STScI and the integration time, we calculate that the scaling factor necessary to remove the continuum contribution to the narrow-band filter using the F555W filter to be about 50 -- 100. Therefore continuum subtraction has only negligible effect on the properties of the extended emission. Unfortunately, we could not empirically measure this scaling factor. This is because the Quad [OII] filters only cover one of the WFCs at a time and in each of the images of the quasars, the stars available for such an estimate were either too faint to be useful, or were too bright and were saturated in the F555W image. To continuum subtract the narrow-band data, we block-averaged the PC continuum images 2 $\times$ 2 to make the scale of the PC continuum images match that of the WFC narrow--band images. The two images were then aligned and the F555W image then scaled by the factor calculated above. The final image is referred to as the Ly$\alpha$ image. The fluxes measured in the Ly$\alpha$ images are in reasonable agreement (always within a factor of 2) with those obtained from the ground by Heckman et al.~(1991a). The number of stars available for the construction of empirical PSF for the narrow-band frames was very limited. However, each individual star used for the construction had many times the counts in the images of the quasars and therefore, even with just a few stars, the signal to noise in the final PSF was sufficient to make a reasonable subtraction of the PSF. We note that in none of the cases did a strong point source appear in the images and thus the method of PSF subtraction was a little bit different from the subtraction of the continuum images. None of the narrow-band images of the quasars had obviously visible diffraction spikes and thus most of the extended emission is little effected by the structure of the PSF. We therefore subtracted until the emission from the central peak roughly blended in with the more extended emission ({\it i.e.,\ } the subtraction did not cause a hole in the position of the nucleus). In Table 2, we quote the results of the PSF subtraction. \subsection{Host Measurements: Other Methods} We attempted several other techniques for estimating the amount of extended emission from each quasar. One relied on scaling the PSF such that its flux in the central 2 pixels matched that of the quasar image. We then took the ratio of the total energy encircled in apertures of increasing radius. This method provides an estimate of the minimum amount of extended flux in each quasar. The other technique we used to make this estimate was to cross-correlate a series of galaxy models plus PSF (to represent the underlying galaxy and quasar) with the image of the quasar. This technique was developed in a series of papers (Phillips \& Davies 1991; Boyce, Phillips, \& Davies 1993). Complete descriptions of the technique and its robustness can be found in the original papers cited above. We made this comparison with a series of model elliptical galaxies of varying half-light radii and ellipticities. Since we and the referee found that given the complexity of the host morphology revealed in the point-spread-function subtracted images, these other methods were perhaps not as convincing as the PSF subtraction. Thus, we will not go into detail of the results of these other methods. It is sufficient to say that they gave estimates very similar to those obtained through PSF subtraction. \section{Results} Contour plots of the HST images of all 5 quasars are shown in Figure 1. To compare the optical and radio structures we also overlaid high resolution radio maps from Lonsdale, Barthel, \& Miley (1993) on the HST images, as shown in Figure 2. For these overlays, the coordinates from the HST images are insufficiently accurate to directly overlay the images using the coordinates given by the standard pipeline reduction. There is an uncertainty of 0.5\arcsec \ to 1\arcsec \ between the absolute HST positions as given by the pipeline reduction and the radio coordinate system. Therefore, we made some assumptions about how to position the radio images relative to the HST images. Using published spectral information available for these quasars, we identified the flattest spectrum component in each radio image, which is presumably associated with the radio AGN, and aligned that component with the most intense pixel ({\it i.e.,\ } the optical quasar nucleus) in the HST image. Second, we checked that our core determined using the method just outlined agreed with the astrometry from Barthel (1984) between high resolution VLA radio maps and the position of the quasar and in every case we found good agreement. \subsection{Notes Individual Objects} \subsubsection{PKS 0445+097} The continuum image of PKS 0445+097 is rather asymmetric and one can discern two components: an asymmetric circum-nuclear host within $\sim$ 1\arcsec \ of the nucleus, elongated along P.A. $\approx$ 240$^\circ$, and a detached ``blob'' at $\sim$ 2\arcsec \ from the nucleus at P.A. $\approx$ 120$^\circ$ The host has a total magnitude of about 22.1$\pm$0.3 magnitudes through the F555W filter. This ``blob'', located about 1.5\arcsec \ ($\sim$17 kpc) to the east-southeast of the quasar nucleus, has a total magnitude of 23.1$\pm$0.2 and is composed of several bright clumps immersed in a more diffuse structure. At surface brightness levels of $\approx$23 m$_{F555W}$ arcsec$^{-2}$, its total extent is approximately 2.5 $\times$ 1.5 arc seconds (along PAs of 35$^\circ \pm$10$^\circ$ and 125$^\circ\pm$20$^\circ$). There is hint of a faint tail of emission leading from the region about the quasar nucleus out toward this blob of emission to the east-southeast. Another interesting feature in the circum-nuclear continuum emission is a $\sim$1\arcsec \ long ``arc'' of emission that curves out eastward of the nucleus and bends to the southeast. The total extent of the circum-nuclear emission, down to surface brightness levels of 23 m$_{F555W}$ arcsec$^{-2}$, is about 2.5 $\times$ 0.8 arc seconds (roughly along PA = 105$^\circ$ and PA = 15$^\circ$; 30 kpc $\times$ 9 kpc for z=2.110). The total extended emission (both the circum-nuclear host and blob to the southeast) comprises about 25\% of the total emission seen from PKS 0445+097 in the F555W filter. The circum-nuclear host contributes about 17\% of the total and the blob to the southeast of the nucleus about 7\%. The F555W continuum image is qualitatively and quantitatively consistent with the broad-band continuum images from Lehnert et al.~(1992). These ground-based images show that the extended emission is blue (consistent with that of a nearby Irregular or Sc galaxy) and extended on spatial scales of about 8 arc seconds ($\sim$100 kpc) to the southeast (Lehnert et al.~1992). The HST image does not show as large of an extent, only about 2 -- 3 arc seconds, but the orientation of the host is similar and the relative fraction of extended emission is the similar (15 -- 40\% in the ground-based data). The HST image is more sensitive to the high surface brightness and more compact structures while the ground-based images with their inferior spatial resolution and larger projected pixel sizes, are more sensitive to the low surface brightness and more diffuse emission. This being the case, inspection of the HST F555W and our previously published ground-based images (Lehnert et al.~1992) suggests that both high surface brightness areas -- the circum-nuclear host and the blob of emission -- are embedded in a diffuse area of emission on scales of tens of kpc preferentially oriented along the axis southeast to northwest but also having some low surface brightness to the north, northeast, and southwest of the nucleus. The narrow-band image of PKS 0445+097 is peculiar. The flux of the brightest source in the image is lower by a factor of $\sim$100 than what we would have expected compared to the flux measured in a ground-based image (Heckman et al.~1991a). The filter we used for the narrow-band imaging of PKS 0445+097 is unusual in one respect compared to the other filters. In the filter holder, the FQUVN-A filter is located within the beam of the planetary camera. To move the filter onto one of the wide-field camera arrays, required the filter holder to be rotated 33$^\circ$. We attribute the discrepancy in the measured flux to an unknown error in the positioning of the filter or of the target and for the remainder of the paper, we will not consider the narrow-band image of PKS 0445+097 further. \subsubsection{MRC 0549-213} The continuum image of MRC 0549-213 reveals a complex structure surrounding the quasar nucleus. The emission is symmetric about the nucleus, with the principal axis of the emission changing from about 90$^\circ$ within a few tenths of an arc second of the position of the nucleus, to a position angle of about 150$^\circ$ at a distance of 1\arcsec . The total magnitude of the quasar (nucleus plus host) is about 19.7 and the magnitude of the host is about 21.3$\pm$0.3 magnitudes in the F555W filter. The fraction of the total brightness contributed by the host is about 23\%. In addition, we see a complex structure of continuum emission about 3.2 arc seconds to the west of the nucleus. This emission region has a curved arc shape and is approximately 1\arcsec \ in size, down to surface brightness levels of $\approx$23 m$_{F555W}$ arcsec$^{-2}$. It has a total magnitude of about 24.0. The F555W continuum image is qualitatively and quantitatively consistent with the U band continuum image in Heckman et al.~(1991). The ground-based U image shows that host is very extended, with the near-nuclear emission (within a few arc seconds of the nucleus) being preferentially oriented along PA $\approx$ 150$^\circ$. Moreover, there is a ``tail'' of emission that extends about 5 arc seconds to the west. This morphology agrees quite well with that seen in our F555W continuum image. Comparing the ground-based and HST image in detail suggests that the emission from the near-nuclear environment of the quasar and the blob to the west must actually be physically connected (deep R-band images of MRC 0549-213 obtained as part of another ground-based program shows a very similar morphology). Our HST image is not sufficiently deep to detect this connection. Also, the ground-based data suggested that about 20\% of the U band flux is extended, consistent with the 23\% estimated from our HST data. The narrow-band image of MRC 0549-213 shows extended structure. Within a few tenths of an arc second of the nucleus, the emission is extended along PA$\approx$ 45$^\circ$. On a scale of a few arc seconds there are two regions of significant emission. One is along PA $\approx$170$^\circ$ and extends out about an arc second from the nucleus. About 1.6 arc second to the east of the quasar nucleus, there is a faint blob of emission that seems to be connected to the quasar nucleus proper. This faint blob of Ly$\alpha$ emission is approximately 1 arc second long in the north-south direction and about 0.5 arc seconds wide in the east-west direction down to our detection limit. The flux from this distinct region of emission is 9.7$\times$ 10$^{-16}$ ergs s$^{-1}$ cm$^{-2}$. In addition, we see several areas of low surface brightness emission near this object. Two regions are particularly noteworthy. One area corresponds to the blob 3.2 arc seconds to the west. There is a $\approx$ 4 $\sigma$ region of Ly$\alpha$ emission over the region of this blob. Also, there is some Ly$\alpha$ emission roughly corresponding to a area of continuum emission about 3 arc seconds away from the nucleus along PA $\approx$ 325$^\circ$. Unfortunately, there does not exist a ground-based Ly$\alpha$ image of this quasar with which to compare. \subsubsection{PKS 1318+113} The continuum image of PKS 1318+113 shows two concentrations of emission, one immediately surrounding the quasar nucleus and the other about 2 arc seconds to the east of the quasar nucleus. Immediately surrounding the quasar nucleus (within about 1\arcsec), the emission is asymmetric, with the brighter isophotes oriented along PA $\approx$ 135$^\circ$ and the fainter isophotes are most extended along PA $\approx$ 180$^\circ$ to 200$^\circ$. The total magnitude of the quasar is about 19.0 in the F555W filter and the host has a magnitude of about 20.1. This implies that the host makes up about 38\% of the total emission from the quasar (nucleus + host). We note that perhaps this is somewhat over-estimated in light of the fact that the cross-correlation technique implies that only about 19\% of the quasar light is extended. The galaxy to the east of the nucleus has a magnitude of about 21.9 (measured in a 2\arcsec \ $\times$ 2\arcsec \ box, which is as large as can be used due to the proximity of this galaxy to the line-of-sight of the quasar). Down to surface brightness levels of $\approx$24 m$_{F555W}$ arcsec$^{-2}$, the extent of the circum nuclear host is about 1 to 1.5 arc seconds. The HST F555W image is similar to the B-band image presented from Heckman et al.~(1991a). The ground-based image shows bright extended emission to the east of the nucleus, with fainter emission to the south and west. The total extent of the ground-based B image is about 6 - 10\arcsec \ from the nucleus. The HST image does not reveal emission quite as extended as this, only about 2 -- 3 arc seconds, but the gross characteristics of the host is similar. The relative fraction of extended emission between the ground-based and HST data are not very similar (16\% in the ground-based B data versus about 38\% in the HST F555W data; although we note that the cross-correlation analysis gives 19\%; see Table 2). This suggests that relatively speaking, the light from the host is concentrated within an 1\arcsec \ of the nucleus (scales not available from the ground). Although again, we note that the cross-correlation analysis gives a result much more consistent with our previous ground-based results. The narrow-band image of PKS 1318+113 shows extended emission (Figure 1). Most of the extended Ly$\alpha$ emission is to the north and east of the nucleus, primarily along PA$\approx$ 45$^\circ$ and is extended on 1\arcsec \ to 2\arcsec \ from the nucleus (down to surface brightnesses of 6.3 $\times$ 10$^{-16}$ ergs s$^{-1}$ cm$^{-2}$ arcsec$^{-2}$. There are several faint regions of emissions within a few arc seconds of the quasar. These regions have fluxes of roughly 1.7 to 7.1 $\times$ 10$^{-16}$ ergs s$^{-1}$ cm$^{-2}$. Moreover, we find reasonable agreement with the morphology of the ground-based Ly$\alpha$ image presented in Heckman et al.~(1991a). In the ground based Ly$\alpha$ images the host was extended along PA$\approx$45$^\circ$ with the most extended emission being on the southwest side of the nucleus. Detailed comparison between the ground-based and HST Ly$\alpha$ images suggest that the highest surface brightness emission is on the northeast side of the nucleus with several bright clumps of the southwest side that is then embedded in a halo of diffuse Ly$\alpha$ emission. One of the most remarkable results of this small HST survey of the host galaxies of high redshift quasars is the interesting spatial relationship between extended Ly$\alpha$ and radio jet emission. In Figure 2, we show an overlay of the Ly$\alpha$ HST image and a VLA A-array map from Lonsdale et al.~(1993). We see that the jet passes between two Ly$\alpha$ emitting blobs southwest of the quasar nucleus. This interaction appears at the point where the jet appears to bend. The two blobs have total fluxes of 1.67$\times$10$^{-16}$ ergs s$^{-1}$ cm$^{-2}$ and 2.01$\times$10$^{-16}$ ergs s cm$^{-2}$ for the eastern-most and western-most emission regions. Measuring the sizes of these blobs, we find that the eastern most blob is approximately circular with a diameter of 0.3\arcsec. The western most blob of the two is approximately 1\arcsec \ $\times$0.35\arcsec \ (long versus short axis oriented PA$\approx$150$^\circ$). We also note that there seems to be another region of Ly$\alpha$ emission along the ``counter-jet'' side of the quasar between the nucleus and the north-eastern radio lobe. \subsubsection{1658+575 (4C 57.29)} The continuum image of 1658+575 (4C 57.29) shows a relatively compact (about 1\arcsec \ across) region of extended emission. The total magnitude of the quasar (nucleus + host) is 18.3 and the magnitude of the host is 20.0. We find that about 21\% of the total emission from the quasar is extended. We see a linear feature along PA$\approx$150$^\circ$ in the extended emission. This feature is very likely to be a residual emission that was not accounted for during PSF subtraction. This is not surprising since 1658+575 is the brightest quasar imaged during this program and hence had the most extended and intense diffraction spikes compared to the other quasar images. Ignoring the linear feature we see that the brightest emission is north and to the southwest of the nucleus. The diameter of the host is only about 1\arcsec \ down to surface brightnesses of 22 m$_{F555W}$ arcsec$^{-2}$. The ground-based B band image of 1658+575 presented in Heckman et al.~(1991a) shows a structure roughly similar to that observed using the HST. The quasar is extended on scales of about 10\arcsec \ in the ground-based image and has a high surface brightness region to the east-southeast of the nucleus with lower surface brightness emission also to the north-northeast and south of the nucleus. The narrow-band image of 1658+575 shows an extended plume of emission to the northeast with some very significant emission also extended to the northwest of the nucleus. There is also a lower surface brightness extension to the south of the nucleus. The most extended emission reaches a radius of about 2\arcsec \ from the nucleus. Comparison with the ground-based Ly$\alpha$ image of Heckman et al.~(1991a) again reveals a very strong similarity between the two images. In the ground-based data Ly$\alpha$ is extended on scales of up to 6\arcsec \ from the nucleus. The most significant of this extended emission is to the northwest through the south side of the nucleus. \subsubsection{PKS 2338+042} The host galaxy of the quasar PKS 2338+042 comprises nearly 40\% of the total continuum emission from the quasar. The host is asymmetric with an ``arm'' of emission that emanates from the nucleus to the south and then bends around to the east. In addition, there is a ``plume'' of emission to the northeast of the nucleus. The total extent of the continuum nebula is about 1.5\arcsec \ down to a surface brightness of 24.0 m$_{F555W}$ arcsec$^{-2}$. The circular contour seen 0.4\arcsec \ to the east of the nucleus in the contour plot is a local minimum in the emission. The 15 GHz radio map of Lonsdale et al.~(1993) shows a ``bent'' core, jet, double lobe source oriented preferentially in an east-west direction. The ``jet'' emanates from the nucleus along PA$\approx$90$^\circ$, with the hotspot of the eastern lobe is at PA$\approx$120$^\circ$. The western hotspot is at PA$\approx$270$^\circ$. Overlaying this map onto the PSF subtracted HST image, we see a close correspondence between features in the HST image and that of the radio image. The maximum intensity seen in the contour plot of the radio emission just to the east of the nucleus along the ``jet'' corresponds to the local minimum we see in the contour plot of the F555W image. This local minimum gives the impression that we are looking down the end of a hollow tube of emission which contains the radio emission. Farther to the east we see that the surface brightness of the continuum emission increases at roughly the same point where the jet seems to ``bend'' towards the hotspot. Moreover, we notice that the rest-frame UV isophotes of the F555W image seem to bend outwards to the west of the nucleus approximately along the same position angle as that to the western radio lobe ({\it i.e.,\ } PA$\approx$270$^\circ$). The F555W continuum image is roughly consistent with the ground-based 4m image taken through the B-filter by Heckman et al.~(1991). This ground-based image shows the most significant emission is to the southeast of the nucleus and is extended about 6\arcsec \ across down to a surface brightness of 27.3 m$_B$ arcsec$^{-2}$. The HST image does not reveal emission quite as extended as this, only a few arc seconds, but the orientation of the host is roughly similar. The HST image however suggests that a much greater fraction of the emission is extended, namely $\sim$40\% versus $\sim$16\% from the ground-based B image. The HST image is obviously more sensitive to the high surface brightness and more compact structures while the ground-based images with their inferior spatial resolution and larger projected pixel sizes, are more sensitive to the low surface brightness and more diffuse emission. This would suggest that perhaps Heckman et al.~(1991) over-subtracted the ground-based image of PKS 2338+042 and that the UV continuum host of PKS 2338+042 is very compact compared to the rest of this small sample. The Ly$\alpha$ image of PKS 2338+042 is also very extended, revealing a host galaxy approximately 2\arcsec \ across down to a surface brightness of 5.4 $\times$ 10$^{-16}$ ergs s$^{-1}$ cm$^{-2}$ arcsec$^{-2}$. The morphology of the host galaxy in Ly$\alpha$ is similar to that of the continuum emission. The basic orientation of the Ly$\alpha$ host is east-west. The most significant piece of the morphology of this image is an ``arm-like'' structure that extends from the nucleus out to about 1 arc second to the east where it terminates in a relatively high. surface brightness region of emission. To the north of this ``arm'' there is a region of relatively low surface brightness compared to its immediate surroundings (a ``hole'' in the extended emission). There are fainter ``plumes'' of emission to the northeast and the south-southeast. At the lowest surface brightness visible in the Ly$\alpha$ image, there is also a faint extension to the northwest. At the lowest surface brightness levels, the orientation of the Ly$\alpha$ is preferentially in the southeast-northwest direction (PA$\approx$ 140$^\circ$) as opposed to the general east-west orientation at higher surface brightnesses. The morphology of the HST Ly$\alpha$ image is similar to that seen in the ground-based Ly$\alpha$ image of Heckman et al.~(1991). The ground-based Ly$\alpha$ image shows a general southeast-northwest orientation with Ly$\alpha$ emission being seen over about 9\arcsec. The ground-based Ly$\alpha$ image is of course a much deeper image than our HST Ly$\alpha$ image, reaching down to a surface brightness of 1.5 $\times$ 10$^{-17}$ ergs s$^{-1}$ cm$^{-2}$ arcsec$^{-2}$. Overlaying the 15 GHz radio image of Lonsdale et al.~(1993) on the HST Ly$\alpha$ image we again see a good correspondence between the highest surface brightness Ly$\alpha$ emission and the radio ``jet''. The Ly$\alpha$ shows a high surface brightness extension about 0.8\arcsec \ to the east of the nucleus. Over this same region the jet of radio emission is observed. It is interesting that over the region of the most intense extended Ly$\alpha$ emission is also the region where we see the ``jet'' of radio emission and where the radio emission undergoes its most severe bending (in projection). Moreover, we again see that to the west of the nucleus, the isophotes ``bend'' outwards from the nucleus over a region about 0.5\arcsec \ out from the nucleus. This is along the same position angle from the nucleus that we see the most distant radio hotspot To elucidate the relationship between the radio and UV continuum and Ly$\alpha$ line emission, we have made a single cut through the 15 GHz radio image from Lonsdale et al.~(1993), and both the PSF subtracted F555W and narrow-band Ly$\alpha$ image (which has not been PSF subtracted). These cuts were made from the highest surface brightness peak in the nucleus of the radio image, in the PSF subtracted F555W, and in the Ly$\alpha$ image and then including all the pixels to the east and west of the nucleus out to a radius of roughly an arc second in both directions. We have normalized and overlayed these cuts in Figure 3. The direction of the cut was selected such that it passed directly along the radio jet that points directly to the east of the nucleus so that we may directly compare the one-dimensional spatially extended radio, UV continuum, and Ly$\alpha$ radial distributions. As can be seen in this Figure, the radio and UV continuum emission are strongly correlated, while the radio and Ly$\alpha$ emission are not. Since this projected cut lies along the direction of the minimum in the spatial distribution of the UV continuum (see Figure 1), the top panel of Figure 3 shows that the jet must pass through the local minimum (best described as a ``hole'') in the distribution of the UV continuum emission. On the other hand, the anti-coincidence of the radio and Ly$\alpha$ distribution is such that at the position where the jet is bending away to the southeast, which is why the radio surface brightness in the bottom panel of Figure 3 is decreasing, the Ly$\alpha$ surface brightness is increasing and reaching a local maximum. The spatial relationship between the Ly$\alpha$ and radio emission is very suggestive of a ``jet-cloud'' interaction in that the area of high surface bright Ly$\alpha$ emission is responsible for bending the jet. While the structure of the Ly$\alpha$ image is rather complex, to aid us in interpreting the data in relationship to a possible ``jet-cloud'' interaction, we wish to estimate the flux from the region of relatively high surface brightness in the Ly$\alpha$ image at the point where the jet bends to the southeast. Isolating the pixels over this region (approximately 0.3\arcsec \ $\times$ 0.3\arcsec \ region about 0.8\arcsec \ from the nucleus), we find a total Ly$\alpha$ flux of 2.1 $\times$10$^{-16}$ ergs s$^{-1}$. \section{Discussion} In this section we discuss the results and their implications for our understanding of the circum-nuclear environments of high redshift quasars. Our sample is too small for a detailed statistical analysis. Thus, we will focus our attention on a few commonalities shared by the quasar hosts and compare these properties with those of high redshift radio galaxies and field galaxies, and low redshift starburst galaxies. \subsection{The Radio-Aligned UV Continuum and Confusion by Intervening Absorber Galaxies} Heckman et al.~(1991a) and Lehnert et al.~(1992), from ground-based images of quasars, found weak evidence for alignment between the rest-frame optical/UV and the radio emission. In a HST/WFPC2 snapshot study of 43 quasars from the 3CR catalog, Lehnert et al.~(1999b) argue that quasars hosts indeed exhibit the ``alignment effect'' in the continuum plus line emission (all of the broad-band HST images in that study have some contribution due to emission lines) but that the effect is slightly weaker than in radio galaxies at similar redshifts. As discussed below, the HST data show that intervening galaxies along the line of sight to the quasar may confuse the morphologies of the hosts: 2 of the 5 quasars from our sample appear to have nearby galaxies seen in projection (PKS 0445+097 and PKS 1318+113). Interestingly, these two quasars were precisely those which showed the greatest mis-alignment between the rest-frame UV continuum host and radio emission in the study of Heckman et al.~(1991a). The evidence that PKS 0445+097 has a nearby intervening system SE of the nucleus is based on Keck 10m spectroscopy (Lehnert \& Becker 1998), as well as morphological and luminosity considerations. The Keck spectrum shows that the SE blob is at a redshift of 0.8384$\pm$0.0002, which is similar to that of the MgII absorption seen against the nuclear continuum of PKS 0445+097 (Barthel, Tytler, \& Thompson 1990). Therefore, it is not surprising that we found no Ly$\alpha$ emission from this ``blob'' at the redshift of the quasar. In addition, Lehnert \& Becker also found that this galaxy is likely to contain a Seyfert 2 nucleus. If we adopt z=0.84 for the redshift of this galaxy, the central wavelength of the F555W filter corresponds to about 2930\AA \ in the rest-frame of the galaxy. This wavelength is close to the wavelengths of the U and B filters and thus extrapolations from flux density measured in the F555W to estimate the flux densities of the U and B filters in the rest-frame of the quasar host are small. Using the spectral energy distribution from Armus et al.~(1997) for this galaxy (approximately that of the a late-type spiral) to extrapolate the measured flux density in the F555W filter to the flux density at the wavelengths of the U and B filters in the rest-frame of the intervening galaxy and correcting for Galactic extinction, we find that the U and B absolute magnitude of the blob to the SE of the nucleus is M$_U$ = $-$21.8 and M$_B$ = $-$21.3. Thus this intervening galaxy is approximately a factor of 2 more luminous than a fiducial Schecter $L^{\ast}$ \ galaxy. Moreover, this intervening galaxy appears to have a very distorted morphology. The HST F555W image shows a galaxy with three knots of emission elongated along PA$\approx$45$^\circ$ with the brightest region not roughly in the center of the emission but towards the northeastern end of the galaxy. The galaxy appears to be nearly edge-on. Comparing the morphology of this galaxy with other intervening absorbers observed with the HST ({\it e.g.,\ } Dickinson \& Steidel 1996; Steidel et al.~1997), we find that the galaxy along the line of sight to PKS 0445+097 is peculiar. Most MgII absorbing galaxy have properties consistent with the general field population of galaxies and hence ``normal'' morphologies and distributions of luminosity similar to field galaxies (Steidel et al.~ 1997; Bergeron \& Boisse~1991; Steidel et al.~1994). However, many of the peculiar morphologies appear to be associated with galaxies that are viewed nearly edge-on (Dickinson \& Steidel 1996). Thus we conclude that even though the morphology appears to be peculiar compared to most MgII absorbing galaxies, its apparently edge-on orientation implies that extinction in the disk may account for its seemingly peculiar morphology. The two color analysis of Armus et al.~(1997) suggests that the color of this galaxy is consistent with Sc spiral galaxy at z=0.84 with about 0.5 magnitudes of addition extinction compared to low redshift Sc spiral galaxies. This additional reddening is consistent with our claim here that extinction may account for this galaxy's seemingly peculiar morphology. However, it could also be that since this galaxy appears to be harboring a Seyfert nucleus (Lehnert \& Becker 1998), it might also be that the complex morphology is associated with a merger event that initiated the Seyfert activity. In the case of PKS 1318+113 no direct spectroscopic evidence exists that its companion to the east is also a foreground object. If this object is at the redshift of the quasar, it would have an implausibly high luminosity ($>$ 25 $L^{\ast}$ and more luminous than the quasar host). There are several other more plausible possibilities for the redshift of this object. The two most plausible redshifts for this object, 0.8388 and 1.0541, which are associated with Mg II absorbers along the line of sight to PKS 1318+113 (Barthel et al.~1990; Jankarkarinen, Hewitt, \& Burbidge 1991). Since this galaxy is bright (about 21.9 in F555W), it is more likely that this galaxy is associated with the MgII absorber at z=0.8388, which is very similar to the case of PKS 0445+097. The absolute magnitude of this galaxy under the same assumptions made previously for the absorber along the line of sight to 0445+097 implies M$_U$ $\approx$$-$23 and M$_B$ $\approx$$-$22.5. These magnitudes are many $L^{\ast}$ and thus we consider associating this galaxy with the MgII absorber at z=0.8388 very implausible and that it is associated with the absorber at z=1.0541 even less likely (see e.g., Bergeron \& Boisse~1991; Steidel et al.~1994). However, such a speculation will have to await spectroscopic observations to determine the redshift of this nearby (in projection) galaxy. \subsection{The Nature of the UV Continuum} We will center our discussion of the UV continuum in the hosts of quasars on two aspects: the origin of radio-aligned UV continuum and the stellar population of the underlying galaxy. \subsubsection{The Radio-Aligned Component} There have been a few hypotheses for the physical causes of extended UV continuum emission in high redshift quasars and radio galaxies. The most viable ones are: 1) star formation stimulated by the radio jet as it propagates outwards from the nucleus (McCarthy et al.~1987; Chambers et al.~1987; De Young 1989; Rees 1989; Begelman \& Cioffi 1989), 2) scattering of light from a hidden quasar by electrons or dust ({\it e.g.,\ } Fabian 1989; Cimatti et al.~1997), 3) inverse Compton scattering of microwave background photons by relativistic electrons in the radio jets or lobes (Daly 1992a; b), and 4) selection effects related to the possible enhancement of radio emission by dense gas which is preferentially located along the galaxy's major axis (Eales 1992). Observing that in fact the radio and UV continuum on small scales are anti-correlated (see also Lehnert 1996), meaning that the high surface brightness radio emission from the jet is actually in a minimum in the rest-frame UV, provides a test of these various proposed schemes. In the model of jet induced star-formation and scattering hypothesis, we might expect to see such anti-correlations on small scales in addition to the ``alignment'' between the radio and UV continuum emission. This might come about for the same reason in both cases. The pressure from the jet would push material outwards both along the jet and perpendicular to it. The high perpendicular pressure might cause clouds to become unstable and collapse and it would also clear material from the region of the jet. In case of the jet-induced star-formation hypothesis, overpressure due to the passing jet might cause these clouds to form stars and in the scattering hypothesis, the over-pressurized clouds would provide for more efficient scattering of the quasar light. In both hypotheses, this would explain the large scale relationship ({\it i.e.,\ } the ``alignment effect'') but on small scales ({\it i.e.,\ } the width of the jet) a anti-coincidence whether generated by star-formation or scattering. However, since it is difficult to understand how the jet can inhibit stars from moving into the regions through which it passes, the jet-induced star-formation scenario only works if the crossing time of the high mass stars is much longer than their evolutionary time scale. Otherwise, the massive stars that are formed at the edge of the jet will fill in the jet region with high surface brightness UV emission. If the time scale for the massive stars to penetrate into the region of the jet is long enough, the massive stars will die out, thus preserving the ``hole'' in the light through which the jet is passing. Interestingly, the radio source in PKS 2338+042 is likely to be young; its observed small radio size (roughly a 10-20 kpc, modulo projection effects) suggests that it is only between 10$^6$ and 10$^7$ years old which is roughly the same order of magnitude as the evolutionary time scale of high mass stars. For example, if the stars are orbiting at 100 km s$^{-1}$, they will transverse 1 kpc (0.1\arcsec \ at the redshift of the quasar in the adopted cosmology) in about 10$^7$ years. The scattering hypothesis does not suffer from such a draw back and is feasible with the only caveat that the jet must be fairly efficient at removing possible scatterers from the regions through which it is passing. Given that the pressure in the jet is estimated to be many orders of magnitude higher than the reasonable pressure of the ISM in a normal galaxy (like the Milky Way for example), such a possibility seems highly plausible. The suggestion of Inverse Compton scattering seems to be ruled out by these observations. Inverse Compton scattering of microwave background photons by relativistic electrons in the radio jets or lobes (Daly 1992a; b), would in fact require that the regions of the highest electron density (likely to be the jets) should have the highest UV surface brightnesses. This is exactly the opposite of what we observed. Assigning the ``alignment effect'' to possible selection effects related to the enhancement of radio emission by dense gas which is preferentially located along the galaxy's major axis (Eales 1992) is an interesting suggestion that seems plausible given the current data set. We have found evidence for ``jet cloud'' interactions in 2 out of the 4 sources imaged at Ly$\alpha$ (see \S 5.3.1). Therefore strong interactions between the radio and ambient interstellar medium are certainly not rare in radio galaxies or quasar hosts (\S 5.3 and references therein). Within this context, it may be that the ``hole'' in the UV continuum may be related to the increased pressure in the region surrounding the jet due to the passage of the jet that in fact prevents it from de-collimating. However, in order to gauge whether or not this speculation is plausible will require more extensive observations. To make this discussion more generic, we note that other quasar hosts appear to have this general alignment, but exhibit detailed spatial anti-coincidence between the radio and rest-frame UV emission (although perhaps not as dramatic as that seen in PKS 2338+042). In HST snapshot data on a large sample of quasars selected from the 3CR sample, Lehnert (1996) found evidence for subtle anti-correlation between radio and rest-frame UV continuum and line emission in these sources (also see de Vries et al.~1996) even though generally the radio emission ``aligned'' with the rest-frame UV continuum and line emission. These anti-coincidences were mainly seen in the sources with complex compact morphologies, roughly similar to the radio morphology of PKS 1318+113 and PKS 2338+042. \subsubsection{A Possible Radio-UV Synchrotron Jet in PKS 0445+097} As has been emphasized previously, an important issue in the study of high redshift radio-loud AGN is the relationship between the relativistic radio-emitting plasma and the line and continuum emission from the host galaxies. To test this hypothesis, we have overlaid a 0.16\arcsec \ resolution image of PKS 0445+097 from Lonsdale et al.~(1993) on the F555W image (Figure 2). Here we have assumed that the bright, compact eastern component is identified with the quasar nucleus on the basis of its inverted radio spectrum ($\alpha^{15GHz}_{5GHz}$=$-$0.6$\pm$0.2; Barthel 1984). The overlay shows that there is a curved optical feature southwest of the quasar which corresponds to the radio jet. The optical and radio flux densities in a 0.5 $\times$ 0.5 arc second area centered on this feature are 0.18$\pm$0.04 $\mu$Jy ($\lambda_{obs}$=5398\AA, the center of the F555W filter) and 13.5$\pm$0.7 mJy ($\lambda_{obs}$=2.0 cm), implying a spectral index (S$_\nu$ $\sim$ $\nu^{-\alpha}$) of $\alpha^O_R$ = 1.1$\pm$0.2. This is consistent with a steepening radio-optical synchrotron spectrum since the radio spectral index of the jet between 5 GHz and 15 GHz is $\alpha^{15GHz}_{5GHz}$=0.8$\pm$0.2 (Barthel 1984) and suggests that the emission may indeed be associated. If this is true, and deep HST imaging polarimetry would be required to prove this, then the optical jet in PKS 0445+097 would be the most luminous and highest redshift jet known (in the rest-frame of the quasar: log L$_B$ = 29.0 ergs s$^{-1}$ Hz$^{-1}$ and log P$_{1.4 GHz}$ = 34.6 ergs s$^{-1}$ Hz$^{-1}$, more than an order-of magnitude more luminous than any previously discovered synchrotron jet; see Dey and van Breugel 1994 for a discussion of known optical/radio synchrotron jets). However, considering the complex and filamentary structure of the circum-nuclear host, there is of course also the possibility that the optical/radio association is accidental, and the $\alpha^O_R$ fortuitously close to the value expected for synchrotron emission. We note also that the 5 GHz map of PKS 0445+097 published by Barthel (1984) shows a small extension northeast from the core, {\it i.e.,\ } opposite to the southwest radio jet, and coincident with the optical extension in that same direction seen in the HST image. No radio spectral index information for this feature is available. \subsubsection{Star Forming Regions in the Quasars Hosts?} If the circum-nuclear UV continuum from the hosts are due to recent star formation, than it is of interest to compare the UV luminosities with low redshift starburst and normal galaxies (Kinney et al.~1993; Donas et al.~1987; Treyer et al. 1998). We find that the typical luminosity of the circum-nuclear host is about 10$^{12}$ $L_{\sun}$ \ at $\approx$1700\AA \ ($\lambda P_\lambda$; Table 3), compared to $\lesssim$ few $\times$ 10$^{11}$ $L_{\sun}$ for normal and starburst galaxies at low redshift (H$_0$=50 km s$^{-1}$ Mpc$^{-1}$). Thus the host galaxies of quasars are at least an order of magnitude more luminous than the most luminous low redshift galaxies in the UV. However, Calzetti, Kinney, \& Storchi-Bergmann (1994) and Meurer et al. (1997) find the typical UV extinction in the Kinney et al.~ sample and starbursts generally to be about 1-3 magnitudes at $\approx$1700\AA. If we correct the most extreme UV-emitting galaxies in the local universe for this amount of UV extinction, they can indeed reach the UV luminosities observed in the circum-nuclear hosts of the quasars. Therefore analogs for the circum-nuclear host associated with these quasars may exist in the local universe, but must be the most extreme UV-luminous galaxies and, moreover, the quasar hosts must be relatively unobscured. There is no good reason to assume why the latter should be the case. To add some further perspective on the nature of the hosts of radio loud quasars, we note that the UV luminosity of the hosts measured in this study are much more luminous than the ``Lyman drop-out'' field galaxies studied by by Steidel and collaborators (Steidel et al.~1996; Giavalisco, Steidel, \& Macchetto 1996). Using the UV (1500\AA) luminosity function for Lyman drop-out galaxies presented by Dickinson (1998), we find that the average UV luminosity of these 5 quasar hosts is about a factor of 10 more luminous than the most luminous Lyman drop-out galaxies. This comparison was made without correcting either sample for the effects of dust extinction. Therefore, it is difficult to associate quasar hosts with the field population of starburst galaxies at high redshift -- even the extremely luminous ones. \subsection{The Extended Ly$\alpha$ Emission} Our quasars typically have Ly$\alpha$ luminosities of $\approx$few $\times$ 10$^{44}$ ergs s$^{-1}$. Ly$\alpha$ luminosities this high are typical of what is observed in high redshift radio galaxies ({\it e.g.,\ } van Oijk et al.~1997; McCarthy 1993). This suggests that quasars and high redshift radio galaxies have similar galaxy-scale environments and ionizing sources, and supports models which attempt to unify radio galaxies and quasars through orientation, evolution, and environment. In addition, even in our relatively short HST exposures, the Ly$\alpha$ emission is extended over tens of kpc; again very similar to what has been observed in high redshift radio galaxies. These results are in agreement with the conclusions of our other investigations of high redshift (z$>$2) quasar hosts ({\it e.g.,\ } Heckman et al.~1991a, b; Lehnert \& Becker 1998). However, quasars offer us an advantage over the radio galaxies. Since we can observe the nucleus more directly, we can estimate the observed ionizing flux emitted by the quasar nucleus. Using the measured UV fluxes from the HST data and using the scaling relations between the flux of UV wavelengths and the total ionizing energy in quasars from Elvis et al.~(1994), we estimate that in order to produce the Ly$\alpha$ emission observed, the host galaxy intercepts only a few percent of the total ionizing luminosity of the quasar. This is not surprising considering the observations of the ``proximity effect'' in the Ly$\alpha$ forest lines. The ``proximity effect'', where the number density of Ly$\alpha$ forest lines is lower near the quasar than away from it ({\it e.g.,\ } Weymann, Carswell, \& Smith 1981; Bechtold 1994), has been interpreted as the hydrogen becoming more highly ionized within the sphere of influence of the quasar. Observing this effect implies that much of the ionizing radiation from quasars must escape to cluster scales and thus the host galaxy of the quasar cannot be completely optically thick to ionizing photons in all directions. This is in agreement with our estimate that the host galaxy and immediate environment of the quasar nucleus only intercepts a few percent of the total ionizing luminosity. Of course it could be less since there could be local sources of ionization such as young stars and/or shocks generated by the mechanical power from the radio jets. To determine the relative importance of local ionization sources would require spatially resolved spectroscopy. The radio/Ly$\alpha$ overlays (Fig. 2) show that there is a good association between the radio emission and the structure of the Ly$\alpha$ emission. First, the principal axis of the radio emission is generally along the same direction as the extended Ly$\alpha$ emission. This is very similar to the radio-aligned extended emission-line regions seen in high redshift radio galaxies ({\it e.g.,\ } McCarthy 1993 and references therein). Second, the surface brightness of the line and radio emission appear anti-correlated to some degree (see {\it e.g.,\ } Fig. 2). This is generally seen at the point where the jet and radio emission are curved. Third, it seems that the brightest Ly$\alpha$ and radio emission are on the same side where the radio lobe is closest to the quasars. This resembles the radio/emission-line asymmetry correlation found for radio galaxies by McCarthy, van Breugel, \& Kapahi (1991). All of these properties may be best understood as being due to strong interaction of the radio sources (jets and lobes) with dense, asymmetrically distributed ambient gas. Numerous examples of this are known in nearby radio galaxies ({\it e.g.,\ } 3C277.3: van Breugel et al.~1985; 3C 405 and 3C 265: Tadhunter 1991; 3C 356: Eales \& Rawlings 1990; PKS 2152-699: Fosbury et al.~1998; PKS 1932-464: Villar-Martin et al.~1998; 3C 171: Clark et al.~1998). \subsubsection{``Jet-Cloud Collisions''} Our observations show two good examples in which jet-cloud collisions seem to occur: PKS 1318+113 and PKS 2338+042. In the case of PKS 1318+113 we observe two emission-line regions between which the radio jet is passing. At this same location the radio jet bends. This radio/optical morphology is very similar to that seen in some nearby radio galaxies, especially Minkowski's Object (van Breugel et al.~ 1985) and is strongly suggestive of a jet cloud interaction. In PKS 2338+042 the spatial resolution is insufficient to allow a similarly detailed examination but the co-spatial bright Ly$\alpha$ and radio knots east of the quasar and the radio jet curvature downstream from this location suggest a similar jet-cloud collision occurs in this object. By analogy to the radio galaxies we will briefly examine whether the observed cloud properties are consistent with such an interpretation. With the limited data in hand (i.e., without high spatial resolution spectroscopy) we can explore only few of the consequences expected from a violent jet-cloud collision. The main questions we can address are 1) are the clouds massive enough to deflect the jets, and 2) can they survive the collisions on time-scales comparable to the radio source ages? The two Ly$\alpha$ blobs SW of the quasar nucleus in PKS 1318+113 have L$_{Ly\alpha}$=1.1$\times$10$^{43}$ ergs s$^{-1}$ and L$_{Ly\alpha}$=1.3$\times$10$^{43}$ ergs s$^{-1}$ to the east and west of the radio emission respectively. Assuming no dust, pure case B recombination for 10,000 K (Osterbrock 1989), cylindrical or spherical symmetry, and using the projected isophotal dimensions apparent in the Ly$\alpha$ image as discussed previously, we then find that $n_e^2f_V$$\sim$0.5 for both blobs, where $n_e$ is the electron density and $f_V$ is the volume filling factor. Since we do not have data to estimate either $n_e$ or $f_V$ independently, we must rely on estimates obtained for other objects. Rough estimates of the volume filling factors for the extended emission line regions in various radio galaxies suggest f$_V$ $\sim$ 10$^{-4}$ to 10$^{-6}$ and $n_e$ $\sim$ 10 - 1000 cm$^{-3}$ (see {\it e.g.,\ } Baum et al.~1992; McCarthy 1993; Lacy \& Rawlings 1994). If for convenience we assume $n_e$ = 100 cm$^{-3}$, then this would imply a volume filling factor of about 5 $\times$ 10$^{-5}$ and thus consistent with values found by previous studies. These estimates would then imply a mass of ionized material in these clouds of about few $\times$ 10$^7$ (n$_e$/100 cm$^{-3}$) (f$_V$/10$^{-4}$) M$_{\sun}$. Making similar assumptions for the Ly$\alpha$ emission-line regions in PKS 2338+042, we find that $n_e^2f_V$$\sim$1 and would thus estimate that the mass of the clouds must be $\approx$ 10$^8$ (n$_e$/100 cm$^{-3}$) (f$_V$/10$^{-4}$) M$_{\sun}$. We note that the above estimates would be similar if we assumed that the gas were shock heated instead of implicitly assuming that the gas is in recombination equilibrium since most of the Hydrogen line emission comes predominately from the post-shock recombination zone ({\it e.g.,\ } Dopita \& Sutherland 1992). Are such masses capable of deflecting the radio jets emanating for the nuclei? Theoretical modeling suggests that jets can be deflected by discrete objects, but only if certain minimal criteria are met. First and foremost, the deflector must be sufficiently massive as so not to be pushed out of the way of the radio jet too quickly. Following the arguments in Icke (1991) and McNamara et al.~(1996), we estimate that deflecting clouds must have a mass, $M_{cloud} \gtrsim 6\times10^4 M_{\sun} ({{0.1 }\over{\eta}})({{L_{rad}} \over{10^{42} ergs \ s^{-1}}})({{c}\over{v_{jet}}})({{1 \ kpc}\over{l_{jet}}})({{t_{jet}} \over {10^6 \ yrs}})^2$, such that it has not moved away by more than the length of the undeflected jet, $l_{jet}$, during the lifetime of the radio source, $t_{jet}$, and where $\eta$ is the conversion efficiency of the jet power, $P_{jet}$, into radio emission such that $L_{rad}=\eta P_{jet}$, and $v_{jet}$ is the velocity of the jet. From the characteristics of the radio emission given in Barthel (1984), we estimate that the radio luminosity of PKS 1318+113 is about 4$\times$10$^{43}$ ergs s$^{-1}$ and of PKS 2338+042, 7$\times$10$^{43}$ ergs s$^{-1}$ and the ages of both sources is likely to be of-order 10$^6$ to 10$^7$ yrs ({{\it e.g.,}} Alexander \& Leahy 1987). Using these estimates we see that the mass necessary to resist being pushed out of the way is about 10$^8$ $M_{\sun}$ (assuming $\eta$ = 0.1, $v_{jet}$ = 0.2c, $l_{jet}$ = 1 kpc, and $t_{jet}$ = 10$^6$ \ yrs). Our rough mass estimates above suggest that the clouds in PKS 1318+113 and PKS 2338+042 are sufficiently high for these clouds to be able to resist the pressure of the radio jets for the relatively short time that they have likely endured the passage of the radio jet. Can the cloud survive the impact from a powerful jet? In both sources we observe what appear to be discrete clouds at the point where the radio jets are deflected. This suggests that a significant portion of the clouds must have survived the impact in order to remain visible; although perhaps the cloud in PKS 1318+113 has been split or had a hole drilled through it. The clouds, in absence of a confining external pressure, will likely be disrupted on a time-scale of the order a few times the sound crossing time. Taking the cloud to have a gas temperature of about 10$^4$ K, we estimate the sound speed $v_s$ = 15 km s$^{-1}$. For cloud sizes of 3-4 kpc (which is the approximate projected sizes of the clouds in our adopted cosmology), we then find that the sound crossing time of these clouds is of-order 10$^8$ yrs. This is much longer than the likely age of the jet, which is probably comparable or smaller than the radio source age which, for powerful double sources is estimated to be of order 10$^6$ to 10$^7$ yrs ({{\it e.g.,}} Alexander \& Leahy 1987). Therefore we find, within the frame work of our assumptions, that there is no particular problem with seeing relatively intact, massive emission-line clouds millions of years after the jet-cloud collisions occurred. The above calculations suggest that 1) quasar hosts may contain rather large amounts of dense clouds and 2) that these can be very efficient at deflecting radio jets during a a significant fraction of the total age of the source ({\it i.e.,\ } roughly 10$^7$ yrs). To make this arguments more general, we note that we only observed good evidence for jet-cloud interaction in two sources, PKS 1318+113 and PKS 2338+042. In two other sources, PKS 1658+575 and MRC 0549-213 we did not see evidence for a jet cloud interaction, and in PKS 0445+097 we suspect that there is something wrong with the narrow-band observation (\S 4.1.1.). Thus, we see strong cloud-jet interaction in two of the four sources. However, PKS 1658+57 and MRC 0549-213 (and also PKS 0445+097) exhibit linear projected radio morphologies suggesting that no dense clouds intercept the jets in these objects; although there is a region of Ly$\alpha$ emission beyond the edge of the eastern radio lobe of MRC 0549-213 perhaps suggesting a large amount of confining material along that direction. Obviously a much larger sample of quasars with high resolution Ly$\alpha$ images are necessary before as statistically significant conclusion can be made. However, our observations strongly suggest that the bent radio structures in radio quasars may very well be due to the interaction of their jets with dense ambient gas (Barthel \& Miley 1988) and that such interactions may be very common and may affect a large percentage of the total radio-loud high redshift quasar population. \subsubsection{Relevance to Unification Schemes} The Ly$\alpha$ and continuum images of PKS 2338+042 and the Ly$\alpha$ image of PKS 1318+113 show relatively obvious signs of interaction between the radio emitting plasma and ambient emission line gas and perhaps even with the stellar population (as probed by the UV continuum emission). These results suggest that the interaction between the radio and the ambient interstellar medium of the host galaxy and cluster-scale environment must be important. In fact, if we take our data literally, they imply that interaction with the ambient medium is important in determining the radio morphology in $\approx$1/2 of the quasars. Clearly, a much larger sample of high redshift quasars need to be observed to determine the exact statistics of the impact of the structure of the host galaxy and cluster-scale environment in influencing the radio emission. However, even for a limited number of quasars, this observation is important. Given that much of the evidence for orientation based unification relies on various aspects of the radio morphology of the sources (linear sizes, lobe arm length asymmetries, predominance of jets, etc) this result allows us to speculate that one would not expect there to be much evidence for orientation-based unification based on radio observations alone. {\bf IF} interactions between the radio and inhomogeneities in the ISM of radio loud objects are important in determining (and limiting) the size and lobe asymmetries in radio sources, then such interaction might dominate over the simple growth of the radio emission with time and orientation effects. Studies that use observations of linear sizes, lobe arm length asymmetries and bends and twists in the radio jets and lobes to test unification schemes (e.g., Gopal-Krishna, Kulkarni, \& Whitta 1996; Kapahi et al. 1995) may in fact get statistically insignificant results not because orientation-based unification is incorrect, but because interacts between the relativistic radio plasma and the ambient ISM and IGM either dominates or provides a significant source of ``noise'' in the observations. This may partially explain why the results of tests of unification schemes using radio data have been so mixed and that orientation-based unification seems most appropriate for a rather limited range of redshifts and generally only for samples of relatively low redshift radio sources (see Barthel 1989 for example). \section{Summary and Conclusions} In this paper, we presented HST WFPC-2 images of spatially-resolved structures (\lq hosts\rq) around five high-redshift radio-loud quasars. The quasars were imaged using the planetary camera with the broad-band F555W (`V') filter and in the wide field camera with a narrow-band filter whose central wavelength is approximately that of redshifted Ly$\alpha$ in each of the quasars. These radio-loud quasars were selected from the earlier imaging survey of quasar \lq\lq host\rq\rq \ by Heckman et al.~(1991a) and Lehnert et al.~(1992). These HST images confirm and extend our earlier ground-based results. From an analysis of the images and a comparison with high resolution VLA radio images we conclude that: \noindent 1) All of the high redshift quasars are extended in both the rest-frame UV continuum and in Ly$\alpha$. We find extended fractions that range from about 5\% to 40\% of the total continuum within a radius of about 1.5\arcsec. In spite of the fact that these images have higher spatial resolutions and relatively short integration times on a small telescope, the morphological agreement between the ground-based images and these HST images is quite good. Moreover, there is reasonable agreement with our estimates of the fraction of the quasar light contributed by the host galaxies in both the HST and ground-based data. Such a result is surprising given the fact that the HST images reveal a wealth of structure within an arc second of the nucleus which is currently unattainable from the ground. \noindent 2) The typical integrated magnitude of the host is V$\sim$22$\pm$0.5. The typical UV luminosity is roughly 10$^{12}$ $L_{\sun}$ ($\lambda P_\lambda$, uncorrected for internal extinction), which is about a factor of 10 higher luminosity than that observed for the ``Lyman drop out'' field galaxies studied by Steidel and collaborators and the most UV luminous zero redshift starburst galaxies. The Ly$\alpha$ images are also spatially-resolved. The typical luminosity of the extended Ly$\alpha$ is about few $\times$ 10$^{44}$ ergs s$^{-1}$. These luminosities require roughly a few percent of the total ionizing radiation of the quasar. \noindent 3) Quasar host generally show ``alignment'' between the radio, Ly$\alpha$, and UV continuum emission. There is clear evidence that the gas ``knows'' about the radio source. This manifests itself in the ``alignment'' between the radio, Ly$\alpha$, and UV continuum emission, in detailed morphological correspondence in some of the sources which suggests ``jet-cloud'' interactions, and in the fact that the brightest radio emission and the side of the radio emission with the shortest projected distance from the nucleus occurs on the same side of the quasar nucleus as the brightest, most significant Ly$\alpha$ emission. These observations of jet-cloud interaction influencing the radio morphologies is a challenge to simple orientation based quasar/radio galaxy unification schemes. This is perhaps why the use of the radio morphology has generally lead to conflicting results when used to judge the appropriateness of orientation based quasar/radio galaxy unification schemes. \noindent 4) The high spatial resolution of the HST has revealed that objects along the line of sight but near the quasars in projection have made a significant contribution to the continuum light from these objects. We note in particular that 0445+097 and 1318+113 were two quasars with strong mis-alignment between the principal emission axes in the radio and the ground-based images at UV rest wavelengths (Heckman et al.~1991a). It is now clear that this mis-alignment was partially due to the contaminating effects of nearby (in projection; less than 2 arc seconds for the quasar nucleus) foreground galaxies. \acknowledgements The authors wish to thank Ray Lucas for his considerable help in making sure that our program went smoothly. Conversations about the complexities of the HST/WFPC2 PSF with Chris Burrows and John Krist were particularly helpful in making the most of the data. We thank the referee, Dr. Eric Smith, whose comments lead to a substantial improvement in the style and presentation of this paper and Dr. Greg Bothun for his conscientious handling of the manuscript in his role as the scientific editor. The work of MDL and WvB at IGPP/LLNL was performed under the auspices of the US Department of Energy under contract W-7405-ENG-48 and the work of MDL at the Sterrewacht Leiden was supported by funds provided by the Dutch Organization for Research (NWO). This work was supported in part by grant number GO-5393 from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. We also acknowledge support from a NATO research grant.
\section{Introduction} \label{L1} Bohr-van Leeuwen's theorem tells us that the orbital magnetism does not appear in classical theory~\cite{Bohr,vanVleck}. The physical argument for this fact is that the diamagnetic current due to cyclotron orbits of electrons in the bulk region is perfectly cancelled by the paramagnetic current due to skipping orbits near the boundary. The orbital magnetism, which is then possible only in quantum mechanics, was originally derived based on the Landau levels~\cite{Landau1}. In the actual derivation, effects of the boundary have not been taken into account explicitly. These effects of boundary on the orbital magnetism have been later investigated by many authors ~\cite{Kubo,DY5,DY6,DY7,DY8,DY9,Robnik,DY12,DY13,NA,DY,Shapiro}. Above all, Kubo~\cite{Kubo} has applied the Wigner representation to an electron system under a magnetic field and shown that, if a confining potential is slowly varying in space compared to the electron wave length, the Landau diamagnetism results. However, this treatment is not valid at low temperature as stressed by Kubo and clear from the expansion parameters. The magnetic moment at $\:T=0\:$, on the other hand, has been shown by Denton~\cite{DY7} and Nemeth~\cite{DY12} to be different from the Landau diamagnetism in a system under a harmonic confining potential. Yoshioka and Fukuyama~\cite{DY} has actually indicated that under a weak field and at low temperature $(\:k_B T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\:$ energy spacing by a confinement), the magnetic moment of the whole system shows a large fluctuation, and as the temperature is raised under such a weak field, fluctuations disappear and Landau diamagnetism is recovered. Hajdu and Shapiro~\cite{Shapiro} pointed out by studying the case of a groove with a width $L$ that the temperature below which such fluctuations appear is such that $k_B T\sim \hbar/\tau_{tr}$ where $\tau_{tr}=L/v_F$, the time of flight for electrons at the Fermi energy across the system. Stimulated by these indications, we will study in this paper the spatial distribution of current in a confined system in order to understand more details of such a variety of orbital magnetism. From the investigations so far, the shape of confining potential, whether it is harmonic or hard wall, is expected not to affect qualitative aspects of the characteristic features of the orbital magnetism. Therefore we assume a harmonic potential which makes the analytical studies possible, and then clarify the relationship between the orbital magnetism and the spatial distribution of current in various regimes of temperature and magnetic field. The organization of this paper is as follows. In \S\ref{L2}, we introduce the model and summarize the general feature of the orbital magnetic moment. In \S\ref{L3}, the spatial distribution of current will be studied in various regions of temperature and magnetic field, and the microscopic conditions for the validity of Landau diamagnetism are clarified. Summary is given in \S\ref{L4}. We take units $k_B=1$ in the following. \section{Magnetic Moment of Two-Dimensional Confined System} \label{L2} In this section, we introduce our model and explain the general properties of the magnetic moment of the system. We consider two-dimensional electrons confined by an isotropic harmonic potential and the magnetic field is applied perpendicular to the system. For simplicity, we consider spinless electrons with the total number $N_0$ and neglect Coulomb interactions between them. We assume $N_0$ is large enough that the difference between a grand canonical ensemble and a canonical one is of no importance and rely on a grand canonical one to derive a magnetic moment. The word ``magnetic moment'' in this paper implies a magnetic moment of the whole system. The Hamiltonian is written as \begin{equation} {\cal H} =\frac{1}{2m}\left(\vecvar{p}+\frac{e}{c}\vecvar{A}\right)^2+ \frac{1}{2}\makebox[0.1em]{}m\makebox[0.1em]{}\omega_0^2\makebox[0.1em]{}\vecvar{r}^2, \end{equation} where $\vecvar{p}$ and $\vecvar{r}$ are two-dimensional vector, $m$ is the electron mass and $(-e)$ is the electron charge. The radius of the system, $R$, is classically defined as \begin{equation} \frac{1}{2}\makebox[0.1em]{}m\makebox[0.1em]{} \omega_0^2\makebox[0.1em]{} R^2=\mu, \label{Ru}\end{equation} where $\mu$ is the chemical potential. By taking a symmetric gauge $\vecvar{A}=\frac{1}{2}\vecvar{H}\times\vecvar{r}$, we can obtain an eigenfunction diagonal with respect to the angular momentum $\alpha$ as \begin{eqnarray} \psi_{n\alpha}(\vecvar{r})&=& \frac{{\rm e}^{{\rm i}\alpha\theta}}{\sqrt{2\pi}} \:R_{n \alpha}(r)\nonumber\\ R_{n \alpha}(r)&=&\frac{1}{l}\sqrt{\frac{n\makebox[0.2em]{}!} {(\makebox[0.2em]{}n+|\alpha|\makebox[0.2em]{})\makebox[0.2em]{}!}} \:\exp\left[-\frac{r^2}{4\makebox[0.1em]{}l^2}\right] \left(\frac{r}{\sqrt{2}\makebox[0.1em]{}l}\right)^{|\alpha|} L_n^{(|\alpha|)}\left[\frac{r^2}{2\makebox[0.1em]{}l^2}\right] \label{WF} \end{eqnarray} where polar coordinates $(\:r,\:\theta\:)$ are used, and $\:n=0,1,2,...$ , $\:\alpha=0, \pm 1, \pm 2,...$ and $l=\sqrt{\hbar/m\omega} \makebox[0.3em]{},\makebox[0.3em]{} \omega=\sqrt{\omega_c^2+(2\makebox[0.1em]{}\omega_0)^2}\makebox[0.3em]{}, \makebox[0.3em]{} \omega_c=eH/mc\:$ being the cyclotron frequency and $L_n^{(\alpha)}$ is the Laguerre polynomial.\\ The eigenenergy of this state, $E_{n \alpha}$, is given by \begin{equation} E_{n\alpha}=\hbar\omega_c\:\frac{\alpha}{2}+\hbar\omega \left(n+\frac{|\alpha|+1}{2}\right). \label{energy} \end{equation} Especially under a extremely strong field $(\omega_c\gg\omega_0)$, this eigenenergy $E_{n\alpha}$ becomes $\hbar\omega_c(n+1/2)$ for negative $\alpha$. So, the quantum number $n$ corresponds to the Landau level index. By use of eq.~(\ref{energy}), the thermodynamic potential $\Omega$ is written as \begin{equation} \Omega=-\frac{1}{\beta}\sum_{n=0}^{\infty}\sum_{\alpha=-\infty}^{\infty} \log\left[\makebox[0.2em]{}1+{\rm e}^{-\beta(E_{n \alpha}-\mu)}\right], \label{Omega} \end{equation} where $\beta=1/T$ and $\mu$ is the chemical potential adjusted to fix the average electron number to $N_0$ at each values of $H$ and $T$.\\ From the thermodynamic potential, the magnetic moment $M$ is given as \begin{eqnarray} M&=&-\left(\frac{\partial\Omega}{\partial H}\right)_{\mu}\nonumber\\ &=&\sum_{n \alpha}\left(-\frac{\partial E_{n\alpha}}{\partial H}\right)f(E_{n\alpha}) \label{Statistic} \end{eqnarray} where $f(E)$ is the Fermi distribution function. Applying the Poisson summation formula~\cite{Landau2} to the sum over $n$ and $\alpha$ in eq.~(\ref{Omega}) (for details, see Appendix~\ref{LA}), one obtains \begin{equation} \Omega=\Omega_0+\Omega_L+\Omega_{osc}\makebox[0.5em]{}, \label{EachOmega} \end{equation} where \begin{eqnarray} \Omega_0&=&-\frac{1}{\beta(\hbar\omega_0)^2} \int_{0}^{\infty}{\rm d}\varepsilon \int_{0}^{\infty}{\rm d}\eta\makebox[0.2em]{} \log\left[\makebox[0.2em]{}1+{\rm e}^{-\beta(\varepsilon+\eta-\mu)}\right] +\frac{1}{12}\makebox[0.2em]{}\mu\makebox[0.5em]{}, \label{Omega0}\\[1.5em] \Omega_L&=&\frac{1}{24}\left(\frac{\omega_c}{\omega_0}\right)^2\mu \makebox[0.5em]{},\\[1.5em] \Omega_{osc}&\simeq&\frac{1}{2\pi\beta}\sum_{k=1}^{\infty}(-1)^k\left[ \left(\frac{\omega}{\omega_0}\right)^2\frac{1}{k^2}-\frac{\pi^2}{3}\right] \frac{\sin\left(2\pi k\mu/\hbar\omega\right)} {\sinh\left(2\pi^2k/\beta\hbar\omega\right)}\nonumber\\ &+&\frac{1}{2\pi\beta}\sum_{\sigma=\pm}\sum_{l=1}^{\infty} \left(\frac{\omega_{\sigma}}{\omega}\right)\frac{1}{l^2} \frac{\sin\left(2\pi l\mu/\hbar\omega_{\sigma}\right)} {\sinh\left(2\pi^2l/\beta\hbar\omega_{\sigma}\right)}\nonumber\\ &+&\frac{1}{\pi\beta}\sum_{\sigma=\pm}\sum_{k=1}^{\infty}\sum_{l=1}^{\infty} \frac{(-1)^k}{l} \left[\makebox[0.5em]{} \frac{\sin\left[\frac{\pi\mu}{\hbar\omega}\left(k-\frac{\omega}{\omega_{\sigma}}l \right)\right] \cos\left[\frac{\pi\mu}{\hbar\omega}\left(k+\frac{\omega}{\omega_{\sigma}} l\right)\right]} {\left(k-\frac{\omega}{\omega_{\sigma}} l\right) \sinh(2\pi^2l/\beta\hbar\omega_{\sigma})}\right.\nonumber \\ &&\makebox[9.2em]{}+ \left.\frac{\sin\left[\frac{\pi\mu}{\hbar\omega}\left(k+\frac{\omega}{\omega_{\sigma}}l \right)\right] \cos\left[\frac{\pi\mu}{\hbar\omega}\left(k-\frac{\omega}{\omega_{\sigma}} l\right)\right]} {\left(k+\frac{\omega}{\omega_{\sigma}} l\right) \sinh(2\pi^2l/\beta\hbar\omega_{\sigma})}\makebox[0.5em]{}\right], \label{osc} \end{eqnarray} where $\omega_{\pm}=(\omega\pm\omega_c)/2$ and the relation $\mu\gg T$ is assumed. In the derivation of oscillatory terms in $\Omega_{osc}$, we used the approximation as follows, \begin{eqnarray} &&\int_{-\beta(\mu-\eta)}^{\infty}{\rm d}\xi\makebox[0.2em]{} \frac{{\rm e}^{\xi}}{({\rm e}^{\xi}+1)^2}\cdot {\rm e}^{{\rm i}\frac{2\pi k}{\beta \hbar\omega}\xi}\nonumber\\ &\simeq& \int_{-\infty}^{\infty}{\rm d}\xi\makebox[0.2em]{} \frac{{\rm e}^{\xi}}{({\rm e}^{\xi}+1)^2}\cdot {\rm e}^{{\rm i}\frac{2\pi k}{\beta \hbar\omega}\xi} \cdot \theta\makebox[0.1em]{}[\makebox[0.1em]{}\mu-\eta\makebox[0.1em]{}], \end{eqnarray} where $ \theta\makebox[0.1em]{}[\makebox[0.1em]{}x\makebox[0.1em]{}]$ is the Heaviside function. As is known from eq.~(\ref{Omega0}), $\Omega_0$ is dependent on a magnetic field only through the chemical potential $\mu(H,\makebox[0.2em]{}T)$, which is regarded as a function of the field and temperature and almost a constant as a function of the field when $\mu/\hbar\omega_c\gg 1$. Therefore the contribution of $\Omega_0$ to the magnetic moment is much smaller than those due to $\Omega_L$ and $\Omega_{osc}$. Based on this result, the field dependence of the magnetic moment is classified into three regions; `` Mesoscopic Fluctuation (MF)'',`` Landau Diamagnetism (LD) '' and `` de Haas-van Alphen (dHvA)'', as is shown in Fig.~\ref{Phase}. ``MF'' corresponds to the region as $\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_{-}\:$, which implies $\:T/\hbar\omega_0 < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}1\:$ under a weak field $(\:\omega_c/\omega_0 < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}1\:)$ and $\makebox[0.3em]{}T/\hbar\omega_0 < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{} (\omega_c/\omega_0)^{-1}$ under a strong field $(\:\omega_c/\omega_0 > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}1\:)$, while ``LD'' corresponds to the region as $T > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_{+}$, which requires $\:T/\hbar\omega_0 > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}1\:$ under a weak field $(\:\omega_c/\omega_0 < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}1\:)$ and $\:T/\hbar\omega_0 > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\omega_c/\omega_0\:$ under a strong field $(\:\omega_c/\omega_0 > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}1\:)$. The other region in Fig.~\ref{Phase} is ``dHvA''. Fig.~\ref{Moment} shows field dependences of magnetic moment at various temperatures; (a),(b) and (c),(d) are respectively under a weak field $(\:\omega_c < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\omega_0\:)$ and a strong field $(\:\omega_c > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\omega_0\:)$. $N_0$ is set at $5000$ in this calculation. At first, we focus on the region of a weak field $(\:\omega_c < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\omega_0\:)$ shown in (a) and (b), where (a) corresponds to ``MF'' and (b) ranges from ``MF'' to ``LD''. At low temperature as $T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\hbar\omega_0$ corresponding to ``MF'', the magnetic moment shows a large fluctuation with respect to the field, as Yoshioka and Fukuyama have shown. In this ``MF'' region, all the oscillatory terms in $\Omega_{osc}$ contribute to the magnetic moment leading to such a large fluctuation, in addition to $\Omega_{L}=-1/2\cdot \chi_{L}H^2$ $(\: \chi_{L}=-1/3\cdot D_0 \mu_{B}^2$, the Landau diamagnetic susceptibility, where $D_0=\mu/(\hbar\omega_0)^2$, the density of states at Fermi energy ), which leads to the usual Landau diamagnetism. Particularly at much lower temperature $(\:T\ll\makebox[0.3em]{}\hbar\omega_0\:)$ and under a much weaker field $(\:\omega_c\ll\omega_0\:)$, the magnetic moment shows a strong paramagnetism, which was first noticed by Meier and Wyder~\cite{DY8} and discussed by Budzin et al.~\cite{DY9} This strong paramagnetism is attributed to the rotational symmetry of the system. Hence, in the presence of weak disorder, a large spatial variation of magnetic moment, either paramagnetic or diamagnetic, is expected, which fact is the cause of a large variance of orbital susceptibility in the limit of weak magnetic field at low temperature \cite{DY17,HF,DY18,DY19,HY}. As the temperature is raised $(\:T > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\hbar\omega_0\:)$ under such a weak field $(\:\omega_c < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\omega_0\:)$, the fluctuation of magnetic moment is reduced and the magnetic moment becomes linearly dependent on the field, the slope of which gives the Landau diamagnetic susceptibilities $\chi_{L}$ corresponding to ``LD''. In this ``LD'' region, the contribution of $\Omega_{osc}$ are reduced and $\Omega_L$ becomes dominant. Fig.~\ref{Moment}(c) and (d) show the field dependence of magnetic moment under a strong field $(\:\omega_c > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\omega_0\:)$ at various temperatures. (c) and (d) range from ``MF'' to ``dHvA'' and from ``dHvA'' to ``LD'', respectively. At low temperature as $T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\hbar\omega_{-}\simeq \hbar\omega_0^2/\omega_c$, magnetic moment shows slow oscillation with a large amplitude and rapid oscillation with a small amplitude, as is shown in Fig.~\ref{Moment}(c). The former oscillation with respect to the field is characterized by $\mu/\hbar\omega\simeq\mu/\hbar\omega_c$, which is caused by the periodic intersections of chemical potential $\mu$ by the Landau level ( the states characterized by $n$ in eq.~(\ref{energy}) ) and corresponds to the de Haas-van Alphen oscillation. The latter oscillation with respect to the field is governed by $\phi/\phi_0$ where $\phi$ is the total flux penetrating the system and $\phi_0=hc/e$, flux quantum. When the total magnetic flux is increased by $\phi_0$, the degeneracy of each Landau level under Fermi energy is increased by unity. This causes the oscillation of the magnetic moment with a period $\phi_0$ as a function of $\phi$ at such a low temperature as $\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\hbar\omega_{-}\simeq \hbar\omega_0^2/\omega_c\:$ which is energy spacing between different angular momentum states at each Landau level. Physically, this oscillation is caused by a coherent motion of electrons along the edge and can be considered as the Aharonov-Bohm oscillation. Such a behavior under a strong field $(\:\omega_c > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\omega_0\:)$ has been noted at $T=0$ by Meir, Wohlman, and Gefen~\cite{NA}. As the temperature is raised so as $\:\hbar\omega_- < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\hbar\omega_+\simeq\hbar\omega_c\:$ under such a strong field $(\:\omega_c > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\omega_0\:)$ corresponding to ``dHvA'', the AB oscillation disappears and there only remains the dHvA oscillation. In this ``dHvA'' region, the first term in $\Omega_{osc}$ has an appreciable contribution in addition to $\Omega_L$. At much higher temperature $(\:T > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\hbar\omega_+\simeq\hbar\omega_c\:)$, the dHvA oscillation disappears and magnetic moment shows a linear field dependence, i.e. the Landau diamagnetism as seen in Fig.~\ref{Moment}(d) e.g. $T/\hbar\omega_0=4$. \section{Spatial Distribution of Current and Magnetic Moment} \label{L3} In this section, we study the relationship between a spatial distribution of current and a magnetic moment of the whole system, the latter of which is given by the thermodynamic potential $\Omega$ in the previous section. The spatial distribution of current in the system is given as follows, \begin{eqnarray} \vecvar{J}(\vecvar{r})&=&{\rm Re} \left< \hat{\psi}^{\dag}(\vecvar{r}) \frac{(-e)}{m}\left( \hat{\vecvar{p}}(\vecvar{r}) +\frac{e}{c}\vecvar{A}(\vecvar{r})\right) \hat{\psi}(\vecvar{r}) \right> \nonumber\\[2mm] &=&J_{\theta}(r)\makebox[0.2em]{}\vecvar{e}_{\theta}, \end{eqnarray} where $<\cdot\cdot\cdot>$ denotes the thermal average and $\hat{\psi}(\vecvar{r})$ is the field operator and $\vecvar{e}_{\theta}=\frac{\partial\vecvar{r}}{\partial\theta}/ |\frac{\partial\vecvar{r}}{\partial\theta}|$ as shown in Fig.~\ref{System}. By use of the eigenstates in eq.~(\ref{WF}) as the basis, $J_{\theta}(r)$ is given as \begin{equation} J_{\theta}(r)=(-e)v_0\sum_{n\alpha}\left[ \alpha\left(\frac{r}{\xi}\right)^{-1}+\frac{\omega_c} {2\makebox[0.1em]{}\omega_0}\left(\frac{r}{\xi}\right) \right]R_{n\alpha}(r)^{2}\makebox[0.3em]{}f(E_{n\alpha}), \label{Jr} \end{equation} where $\xi=\sqrt{\hbar/m\makebox[0.1em]{}\omega_0}\makebox[0.3em]{}$, the characteristic length, and $v_0=\omega_0\makebox[0.1em]{}\xi\makebox[0.3em]{}$, the characteristic velocity of electrons.\\ This local current $\vecvar{J}(\vecvar{r})$ induces a magnetic moment $\vecvar{M}(\vecvar{r})$ given as follows, \begin{eqnarray} \vecvar{M}(\vecvar{r})&=&\frac{1}{2c}\vecvar{r}\times\vecvar{J}(\vecvar{r})\nonumber\\ &=&\frac{1}{2c}\makebox[0.2em]{}r\makebox[0.2em]{}J_{\theta}(r)\makebox[0.2em]{} \vecvar{e}_{z}\nonumber\\[2mm] &\equiv& M_z(r)\makebox[0.2em]{}\vecvar{e}_{z}. \end{eqnarray} Therefore, by integrating this magnetic moment with respect to $\vecvar{r}$, we get the relation between the total magnetic moment $M$ ($z$-component) and the local current $J_{\theta}(r)$ as \begin{eqnarray} M&=&\int{\rm d}S\makebox[0.3em]{} M_{z}(r)\nonumber\\ &=&\frac{\pi}{c}\int_{0}^{\infty}{\rm d}r\makebox[0.3em]{} r^2 J_{\theta}(r). \label{MJ} \end{eqnarray} This expression of magnetic moment derived from the local current density coincides with the one in eq.~(\ref{Statistic}) derived from the thermodynamic potential (see Appendix~\ref{LB}). \subsection{Region of Weak Magnetic Field $(\:\omega_c < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\: \omega_0\:)$} Here, we focus on the properties of the spatial distribution of current in the weak field region $(\:\omega_c < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\omega_0\:)$. At low temperature $(\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_0$, i.e. in ``MF''), the magnetic moment shows a large fluctuation as a function of the field. The spatial distribution of current in such a situation is shown in Fig.~\ref{Current}(a), which indicates that $J_{\theta}(r)$ can be either positive or negative, i.e. paramagnetic or diamagnetic, respectively. In the bulk region ( mainly, $\:r < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}R\:)$, large currents flow paramagnetically or diamagnetically depending on $r$. This large bulk currents are sensitive to the strength of the field and lead to the large fluctuation of magnetic moment. This behavior of $J_{\theta}(r)$ is a characteristic feature of local currents in the region of ``MF''. As the temperature is raised $(\:T > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_0\:)$ under such a weak field, the fluctuation of magnetic moment disappears and the magnetic moment shows the Landau diamagnetism as discussed in the previous section. According with this change of magnetic moment, the spatial distribution of current $J_{\theta}(r)$ is changed as shown in Fig.~\ref{Current}(b); the fluctuating large bulk currents are immediately reduced and finally the diamagnetic current flowing along the edge ($\:r\simeq R\:$) only survives. It is seen that this property of the current distribution in ``LD'' is characteristic of two-dimensional confined system as deduced from the Kubo's formula of the current distribution, which will be explained in the following. Based on the Wigner representation, Kubo~\cite{Kubo} derived the analytic form of a current distribution proportional to a magnetic field and leading to the Landau diamagnetism in the system under a confining potential $V(r)$, which is assumed to be slowly varying in space compared to the electron wave length. The expansion parameters in this theory are $\:\hbar^2 e \makebox[0.1em]{}H\makebox[0.1em]{} \frac{{\rm d}V(r)}{{\rm d}r} /m^{3/2}c\makebox[0.2em]{} T^{5/2}\:$, $\:\:(\hbar e H/m\makebox[0.1em]{} c\makebox[0.1em]{} T)^2\:$, $\:\:\hbar^{2}\left(\frac{{\rm d}V(r)}{{\rm d}r}\right)^{2} /m\makebox[0.1em]{} T^{3}\:$ and $\:\hbar^{2}\makebox[0.1em]{}\frac{{\rm d}^{2}V(r)}{{\rm d}r^{2}} /m\makebox[0.1em]{} T^{2}\:$. In the present model of harmonic confining potential, the parameters are explicitly given as $\:\:\omega_c/\omega_0\cdot(T/\hbar\omega_0)^{-5/2} \cdot R/\xi\:$, $\:\:(\omega_c/\omega_0)^2\cdot(T/\hbar\omega_0)^{-2}\:$, $\:\:(T/\hbar\omega_0)^{-3}\cdot(R/\xi)^{2}\:$ and $\:(T/\hbar\omega_0)^{-2}\:$ respectively, by replacing $r$ with the system radius $R$. After all, $\:T/\hbar\omega_0 > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:(R/\xi)^{2/3}\:$ and $\:T/\hbar\omega_0 > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\omega_c/\omega_0\:$ are required for the validity of the expansion in the Wigner representation, which are actually a part of the region ``LD'' in Fig.~\ref{Phase} ( However, such an expansion turns out to be an asymptotic one since our analytic result shows that $\hbar=0$ is an essential singularity. See Appendix~\ref{Relation} for a detail. ). The spatial distribution of current, which is valid under such a condition, is given as follows in general, \begin{equation} \vecvar{J}(\vecvar{r})= \frac{1}{3}\makebox[0.2em]{}c\makebox[0.1em]{}\mu_B^2 \makebox[0.2em]{} \vecvar{H}\times\nabla n(\vecvar{r}), \end{equation} where \begin{equation} n(\vecvar{r})=\frac{1}{h^d}\int{\rm d}\vecvar{\pi}\int{\rm d}E \makebox[0.3em]{}\delta \makebox[0.1em]{}' \left(E-\frac{\vecvar{\pi}^2}{2m}-V(\vecvar{r})\right)f(E), \label{integral} \end{equation} $d$ is the dimension of the system and $\vecvar{\pi}$ corresponds to a physical momentum. The magnetic moment due to this current density becomes $\chi_L H$, the Landau diamagnetism. In the present two-dimensional system, $\vecvar{\pi}$ and $\makebox[0.2em]{} E$-integration in eq.~(\ref{integral}) can be easily performed and $n(\vecvar{r})$ is given by \begin{equation} n(\vecvar{r})=\frac{2\pi m}{\hbar^2}f(V(\vecvar{r})), \end{equation} and $\vecvar{J}(\vecvar{r})$ becomes \begin{equation} \vecvar{J}(\vecvar{r})= \frac{1}{3}\makebox[0.2em]{}c\makebox[0.1em]{}\mu_B^2 \cdot \frac{2\pi m }{\hbar^2}\cdot \vecvar{H}\times\nabla V(\vecvar{r})\cdot f\makebox[0.1em]{}'(V(\vecvar{r})). \end{equation} Due to the factor $f\makebox[0.1em]{}'(V(\vecvar{r}))$ which reflects Fermi degeneracy, this form of a current distribution means that the diamagnetic current flows only in the region $V(\vecvar{r})\simeq \mu\makebox[0.2em]{}$, i.e. the edge of the system, even if the potential is varying spatially in the bulk region.\\ Applying the above formula to the present model $V(r)=m\makebox[0.1em]{}\omega_0^2\makebox[0.1em]{} r^2/\makebox[0.1em]{}2$, we get $\vecvar{J}(\vecvar{r})$ as \begin{equation} \vecvar{J}(\vecvar{r})=-\makebox[0.2em]{} \frac{j_0}{6}\cdot\frac{\omega_c}{\omega_0}\cdot \left(\frac{T}{\hbar\omega_0}\right)^{-1}\cdot\frac{r}{\xi}\cdot \frac{{\rm e}^{\beta(V(r)-\mu)}} {(\makebox[0.2em]{}{\rm e}^{\beta(V(r)-\mu)}+1\makebox[0.2em]{})^2}\cdot \vecvar{e}_{\theta}, \label{Wigner} \end{equation} where \begin{equation} \beta(V(r)-\mu)= \frac{1}{2}\cdot\left(\frac{T}{\hbar\omega_0}\right)^{-1} \left[\left(\frac{r}{\xi}\right)^2-\left(\frac{R}{\xi}\right)^2\right], \end{equation} and $j_0=e\makebox[0.1em]{}\omega_0/4\pi\makebox[0.1em]{}\xi$ the characteristic current density. The current distribution calculated by eq.~(\ref{Wigner}) coincides with the one obtained from eq.~(\ref{Jr}). ( For example, at $\omega_c/\omega_0=0.1$ and $T/\hbar\omega_0=10$, the agreement is within $0.5\%$ numerically.) In comparison, the case of $d=3$ which Kubo considered appears somewhat different, since the current both at the surface region ( the region as $\:V(\vecvar{r})\simeq\mu\:)$ and the bulk region can contribute to the magnetic moment. To see this, we approximate a Fermi distribution function in eq.~(\ref{integral}) by the linearized form around the Fermi energy as \begin{eqnarray} f(E)\simeq \left\{ \begin{array}{ll} \makebox[2em]{} 1 \makebox[3.8em]{} (\makebox[0.5em]{}E<\mu-2\makebox[0.1em]{}T\makebox[0.5em]{})\\ \makebox[0.3em]{} \frac{1}{2}-\frac{E-\mu}{4\makebox[0.1em]{}T}\makebox[2.2em]{} (\makebox[0.5em]{}|\makebox[0.1em]{}E-\mu\makebox[0.1em]{}|\leq 2 \makebox[0.1em]{}T\makebox[0.5em]{})\makebox[1em]{}.\\ \makebox[1.8em]{}0\makebox[4em]{} (\makebox[0.5em]{}E>\mu+2\makebox[0.1em]{}T\makebox[0.5em]{}) \end{array}\right. \end{eqnarray} We assume the confining potential $V(\vecvar{r})$ is only dependent on $r$ in cylindrical coordinates $(r,\theta,z)$. Then the current distribution is given by \begin{equation} \vecvar{J}(\vecvar{r})=-\frac{\gamma H}{\sqrt{T}}\cdot V'(r)\cdot g(V(r))\makebox[0.3em]{}\vecvar{e}_{\theta}, \end{equation} where $\gamma=\sqrt{2}\pi m^{3/2}c\mu_B^{2}/3h^{3}$ and $g(V(r))$, the function corresponding to $f'(V(r))$ in the case of $d=2$ is given as \begin{eqnarray} g(V(r))= \left\{ \begin{array}{ll} \makebox[0em]{} \sqrt{\frac{\mu-V(r)}{T}+2}-\sqrt{\frac{\mu-V(r)}{T}-2} \makebox[2.8em]{} (\makebox[0.5em]{}V(r)<\mu-2\makebox[0.1em]{}T\makebox[0.5em]{})\\[0.5em] \makebox[3.5em]{} \sqrt{\frac{\mu-V(r)}{T}+2} \makebox[6em]{} (\makebox[0.5em]{}|\makebox[0.2em]{}V(r)-\mu\makebox[0.2em]{}|\leq 2 \makebox[0.1em]{}T\makebox[0.5em]{}) \makebox[1em]{}.\\[0.5em] \makebox[6.5em]{}0\makebox[8.1em]{} (\makebox[0.5em]{}V(r)>\mu+2\makebox[0.1em]{}T\makebox[0.5em]{}) \end{array}\right. \end{eqnarray} In the bulk region where $\:V(r)\ll \mu-2\makebox[0.1em]{}T\:$, $\:g(V(r))$ becomes \begin{equation} g(V(r))\simeq \frac{2}{\sqrt{\frac{\mu-V(r)}{T}}}\ll 1\makebox[1em]{}, \end{equation} while at the surface region, $|\makebox[0.1em]{}V(r)-\mu\makebox[0.1em]{}|<2\makebox[0.1em]{}T\:$, $\:g(V(r))\sim 1\:$. Therefore, it can be said that the current causing the Landau diamagnetism is mainly induced at the surface also in a three-dimensional system, which is not so clear as the case of $\:d=2$. At low temperature as $\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_0\:$, on the other hand, the contributions of higher order terms in magnetic field become larger and the above formula of current density is not valid. Then, the current distribution changes dramatically at around the temperature $T=\hbar\omega_0$ under a weak field $(\:\omega_c < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\omega_0\:)$ as is shown in Fig.~\ref{Current}(a) and (b) and this change is clearly reflected in a magnetic moment of the system. Hajdu and Shapiro~\cite{Shapiro} studied a two-dimensional system under a confining potential $V(\vecvar{r})=m\omega_0^2 x^2/2$ ( i.e. harmonic groove ) with a width $L_x$ and an arbitrary long length $L_y$ by imposing a periodic boundary condition in the $y$-direction, and pointed out that the temperature as $T=\hbar\omega_0$ below which magnetic moment shows a large fluctuation under a weak field $(\:\omega_c < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\omega_0\:)$ corresponds to $\hbar/\tau_{tr}$ where $\tau_{tr}=L_x/v_F$, time of propagation for electrons at the Fermi energy across the groove. The physical reason why the field dependence of magnetic moment shows such a dramatic difference depending on the temperature is understood as follow. When the effect of a confining potential $V(\vecvar{r})$ is considered perturbative, the thermal Green's function of a degenerate electrons has the following damping factor, \begin{equation} G(x,\tau=0)\makebox[0.3em]{} \propto \makebox[0.5em]{} {\rm e}^{-\frac{\pi T}{\hbar v_F}x}. \end{equation} Therefore, the length $\hbar v_F/\pi T\equiv l_c$ is regarded as the coherence length of degenerate electrons. Under the condition $l_c > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{} L$, where $L$ is a system length, electrons near the Fermi surface can propagate from one side of the system to the other side with a small damping and experience the multiple reflection by the boundary potential as shown in Fig.~{\ref{reflection}}(a). Therefore electrons near the Fermi surface, which play an essential role in the orbital magnetism of degenerate electron system, are strongly affected by the boundary potential. In this model, the condition $l_c > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{} L\makebox[0.3em]{}$ corresponds to $\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_0\makebox[0.3em]{}$ i.e. ``MF''. As shown in Fig.~\ref{Current} (a), it is seen that this multiple reflection induces large currents irregularly distributed paramagnetically or diamagnetically in the bulk region, which are sensitive to the strength of a magnetic field and causes the large fluctuation of a magnetic moment as a function of the field. On the other hand, under the condition $l_c < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{} L\makebox[0.3em]{}$ $(\:T > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_0\:)$, the effect of the multiple reflection by the boundary potential wall is reduced as shown in Fig.~\ref{reflection}(b), and this is considered to lead to the suppression of the bulk currents and the recovery of the Landau diamagnetism. Considering the case of a harmonic groove system based on this idea, which Hajdu and Shapiro~\cite{Shapiro} studied, it is natural that the relative magnitude of $l_c$ and the width $L_x$ affects the field dependence of magnetic moment while $L_y$ does not affect it qualitatively, since the $x$-direction is actually confined by a harmonic potential and a multiple reflection can happen only in the $x$-direction. Robnik~\cite{Robnik} discussed the size effect on the zero-field susceptibility in a system confined by a hard wall, based on the Green's function method, and concluded that at $T=0$ the contribution of the boundary wall to the susceptibility is always paramagnetic but with the order of magnitude $(k_F L)^{-1}$ compared to the Landau susceptibility, where $k_F$ is a Fermi wave number and $L$ is a system length. As is clear from Fig.~\ref{Moment}(a), this evaluation of a size effect due to the confining potential is too small. This disagreement can be attributed to the fact that the multiple reflection by the boundary potential wall is disregarded in ref. 9. \subsection{Region of Strong Magnetic Field $(\:\omega_c > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\:\omega_0\:)$} Under a strong magnetic field $(\makebox[0.3em]{}\omega_c > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{} \omega_0\makebox[0.3em]{})$, the energy spectrum in eq.~(\ref{energy}) becomes close to the one without the confining potential, i.e. the Landau level ( the states characterized by $n$ in eq.~(\ref{energy}) ). At low temperature as $T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_c$, magnetic moment is affected by this Landau quantization and oscillates as a function of the field as is shown in Fig.~\ref{Moment}(c) and (d) (This is the familiar de Haas-van Alphen effect). In this ``dHvA'', the spatial distribution of current is as is shown in Fig.~\ref{Current}(c). Several diamagnetic peaks can be seen and paramagnetic currents flow between them. These bulk currents are much larger than the edge current in ``LD'', although the diamagnetic currents are almost cancelled by the paramagnetic ones as an average. Each diamagnetic peak comes from the state with the same $n$ in eq.~(\ref{energy}) which corresponds to Landau level. In this model, each Landau level has a different degeneracy with respect to the angular momentum, $\alpha$, due to the confining potential and the lower Landau level has the higher degeneracy. Therefore, the spatial extent of the lower Landau level with such a degeneracy becomes larger and is reflected in the current distribution as the spatially separated diamagnetic peaks. As the field is raised, the number of Landau level below the Fermi energy decreases and the diamagnetic peaks closest to the center of the system disappear one by one. This change of the current distribution causes the large oscillation of the magnetic moment as a function of the field. At much lower temperature in ``MF'' $(\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_{-}\:)$ where the AB effect appears, the current distribution is almost same as in ``dHvA'' region. At higher temperature as $T > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_c$, the diamagnetic peaks in the bulk region are smeared out and there remain only edge current as is shown in Fig.~\ref{Current}(d). This edge current is spatially broadened compared to the one at lower temperature in ``LD'' shown in Fig.~\ref{Current}(b). This temperature dependence of edge current is understood from eq.~(\ref{Wigner}) where the spatial extent of edge current corresponds to $|\makebox[0.1em]{}V(r)-\mu\makebox[0.1em]{}| < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}T$. \section{Summary} \label{L4} We have studied the relation between the spatial distribution of current and the magnetic moment of the whole system in a two-dimensional electron system under an isotropic harmonic potential as $\:V(\vecvar{r})=m\makebox[0.1em]{}\omega_0^2\makebox[0.1em]{}r^2/\makebox[0.1em]{}2\:$. It is found that characteristic dependences of magnetic moment on temperature and magnetic field are clearly reflected in the spatial distribution of current. Under a weak field $(\:\omega_c < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\omega_0\:)$, the field dependence of magnetic moment dramatically changes at around the temperature $\:T=\hbar\omega_0\:$. In the low temperature region $(\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_0\:)$, magnetic moment shows a large fluctuation as a function of the field, as was indicated by Yoshioka and Fukuyama~\cite{DY}. We attribute this large fluctuation of magnetic moment to a multiple reflection by the boundary confining potential, which becomes important once the coherence length of degenerate electrons $\:l_c=\hbar v_F/\pi T\:$ gets longer than a system length. In the present model, the system length is characterized by the radius $R$ and the above condition leads to $l_c > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}R$ implying $\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_0\:$. At such low temperature, it is seen that the multiple reflection induces large currents irregularly distributed paramagnetically or diamagnetically in the bulk region, which are sensitive to the strength of the field and cause a large fluctuation of the magnetic moment. As the temperature is raised $(\:T > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_0\:)$, the fluctuations of the magnetic moment are reduced and the usual Landau diamagnetism is recovered. Corresponding to this change, it is found that the large currents in the bulk region are immediately reduced and finally the diamagnetic current flowing along the edge ( in the region satisfying $V(\vecvar{r})\simeq\mu$ ) only survives, which leads to the Landau diamagnetism. This edge current is regarded characteristic of a two-dimensional confined system, as inferred from the Kubo~\cite{Kubo}'s formula of the diamagnetic current distribution which is derived for high temperature as $\:T > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_0\cdot (R/\xi)^{2/3}\:$ and $\:\hbar\omega_c\:$, where $\:\xi=\sqrt{\hbar/m\omega_0}\:$, the characteristic length. From this formula, the cancellation of the bulk currents is seen to be due to the Fermi degeneracy. It is also noted that the persistence of the diamagnetic current at the surface is also seen in a three-dimensional system. Under a strong field $(\:\omega_c > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\omega_0\:)$, dHvA effect appears at low temperature as $\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_c\:$. In this situation, the current is distributed in the bulk region and diamagnetic peaks appear, although paramagnetic currents flow between the peaks, which almost cancel the contribution of the diamagnetic peaks as an average. Here, each peak is due to the Landau level with different degeneracy because of the confining potential. As the field is increased, this diamagnetic peak disappears one by one corresponding to the decrease of the number of Landau levels under the Fermi energy, and this causes the large oscillation of magnetic moment (dHvA effect). At much lower temperature $(\:T < \kern -11.8pt \lower 5pt \hbox{$\displaystyle \sim$}\makebox[0.3em]{}\hbar\omega_{-} \simeq\makebox[0.3em]{}\hbar\omega_0^2/\omega_c\:$, the energy spacing between different angular momentum states at each Landau level ), the small but rapid oscillation with a period $\phi_0=hc/e$ appears as a function of the total flux $\phi$ in addition to the dHvA oscillation, , which was originally found by Meir et al~\cite{NA} at $T=0$. This oscillation is caused by a coherent motion of electrons along the edge and can be called AB oscillation. On the other hand, at high temperature as $T > \kern -12pt \lower 5pt \hbox{$\displaystyle \sim$}\:\:\hbar\omega_c$ under such a strong field, the bulk currents are quickly reduced and the Landau diamagnetism by the edge current is recovered as in the case under a weak field. It should be noticed that this edge current does not cause AB oscillations, because the current is due to incoherent motions of electrons. \section*{Acknowledgment} One of the authors (Y. I) would like to express his sincerest gratitude to Hiroshi Kohno for stimulating and instructive discussions and thank Masakazu Murakami for instructive and useful discussions.
\section*{Acknowledgements} This work was supported by Grand-in-Aid for encouragement of Young Scientists from the Ministry of Education, Science and Culture of Japan and also by CREST from the Japan Science and Technology Corporation.
\section{Introduction} It is widely accepted that polymers represent complicated structures which consist of coupled one-dimensional (1D) chains, and many unusual properties can be understood based on their quasi-1D feature. During last decades analytical and numerical studies revealed that the phenomenon of Anderson localization is a characteristic feature of 1D systems. For a quasi-1D system the interchain coupling tends to delocalize the electronic states, which may lead to extended states when interchain diffusion rate $\omega$ exceeds a threshold value $\omega_{c}$\cite{s1}. Moreover, conductance of several polymers, such as stretched polyacetyllene and polyanilline, is enhanced by a few orders of magnitude upon doping and can even reach high values typical of metals, while the temperature dependence of static conductance ($g_{dc}$) widely vary from sample to sample\cite{s2,s3,s4,s5,s6,s7,s8}. In order to understand the effect of chain windings and random interchain couplings on the electronic transport properties in quasi-1D polymers, Xiong {\it et al} \cite{s10,s100} have investigated the transmission coefficient and conductance for models of winding and randomly coupled polymers. It is proved that for such systems there exist conducting subbands with ``dilute'' structure which can account for the high values of conductance and its insulator-like temperature dependence of some polymers. On the other hand, Samukhin {\it et. al} \cite{s9} suggested that the peculiar features of conducting polymers can be derived based on a special hierarchical structure--a nearly 1D fractal. In this paper we investigate the transmission spectrum of the hierarchical structure proposed in \cite{s9}. We prove that this fractal network is equivalent to a 1D chain with hierarchical aperiodic structure for the transport of electrons in the stretched direction. By solving the Schr{\"o}dinger equation of such an effective 1D model by the standard transfer matrix technique, we obtain the transmission coefficient as a function of energy for different stages (or size) of the structure. We find that there also exist ``dilute'' conducting subbands as those in Ref. \cite{s10} and \cite{s100}, but the structure in these subbands exhibits some kind of self-similarity and become more complicated when the stage of the system increases. As a result, the calculated conductance sensitively depends on the structure parameters $(m,n)$ of the original hierarchical fractal and on the position of the Fermi level. The paper is organized as follows: In the next section we present the derivation of the equivalent 1D model and the basic formalism. In Section III the calculated transmission spectrum and conductance are illustrated. Section IV is devoted to a brief summary and discussion. \section{Basic formalism} According to Samukhin {\it et. al}, the hierarchical model for the stretched polymer is constructed in the following way: as a first stage, $n$ primitive bonds are taken to form an $n$-length chain along stretched direction, and $m$ threads of such chains are combined at both ends to form an $(m,n)$ boundle with two common terminals; in the following stages the boundles formed in the previous stage are taken as the elementary bonds and the same procedure is performed to form new boundles. We are interested in the variation of the transmission spectrum in increasing the number of stages $N$, i.e. the size of the system. In measuring the conductance one needs to link electrodes to both ends, (see Fig.1). We use two integers, $(i,j)$, to label the sites, where $i$ indicates the position in the longitudinal direction and $j$ in the transverse direction. If the number of stages of the system is $N$, the longitudinal index $i$ is varying in the range of $[0,n^N]$. For a given $i$, the number of sites in the transverse direction is \begin{equation} \label{rho_N} \rho_N(i)=\left\{ \begin {array}{l} m^N , \text{ if } \text{mod}(i,n)\neq 0 ;\\ m^{N-\alpha_N(i) } \text{ if } \text{mod}(i,n^{\alpha_N(i)}) =0 \text{ and } \text{mod}(i,n^{\alpha_N(i) +1})\neq 0, \alpha_N (i) \text{ is integer in }(0,N); \\ 1, \text{ if } i\leq 0 \text{ or } i\geq n^N, \end{array} \right. \end{equation} where $\text{mod}(x,y)$ denotes the remainder on the division of $x$ by $y$. Thus, in the site index $(i,j)$, $j$ is ranging in $[1,\rho_N(i)]$. From the tight binding approach, the Schr\"{o}dinger equation can be written as \begin{equation} \label{ham} E\psi_{i,j}= \epsilon_{i,j}\psi_{i,j}+\sum_{\delta (i,j)}t_{ij,\delta (i,j)} \psi_{\delta (i,j)}, \end{equation} where $\psi_{i,j}$ is the coefficient of the wave function on site $(i,j)$, $\delta (i,j)$ denotes the position of next neighbors of $(i,j)$, $\epsilon_{i,j}$ is its site energy, $t_{ij,\delta(i,j)}$ the hopping between sites $(i,j)$ and $\delta (i,j)$. As we are only interested in the topological structure of the system, we can assume that the site energies and the nearest-neighbor hoppings are independent of the position. In this case one can set the site energy to be the energy origin and the hopping to be the energy unit. By summing over the transverse sites for a given $i$, Eq. (\ref{ham}) becomes \begin{equation} \label{1d} E\psi_i=\beta_N(i,i+1)\psi_{i+1}+\beta_N(i,i-1)\psi_{i-1}, \end{equation} where \[ \psi_i=\frac{1}{\sqrt{\rho_N(i)}}\sum_{j=1}^{\rho_N(i)}\psi_{i,j}, \] and \[ \beta_N(i,i')=\max \left( \sqrt{\frac{\rho_N(i)}{\rho_N(i')}}, \sqrt{\frac{\rho_N(i')}{\rho_N(i)}}\right). \] From Eq. (\ref{rho_N}) it is easy to see that \begin{equation} \beta_N(i,i')=m^{|\alpha_N(i)-\alpha_N(i')|/2}, \end{equation} where $\alpha_N(i)$ is the largest power of factor $n^{\alpha_N(i)}$ of integer $i$ in the range $[1,n^N-1]$, as defined in Eq. (\ref{rho_N}), and we set $\alpha_N(i)=N$ for $i\leq 0$ and $i\geq n^N$. From Eq. (\ref{1d}) one can see that the fractal structure reduces to a 1D effective Hamiltonian with special aperiodic structure. In the derivation we have adopted the summation over the transverse sites which erases the degree of freedom of electrons in this direction. This has no effect on the longitudinal transport properties for the one-channel incoming and outgoing leads of the present geometry. The situation becomes different if multichannels are connected to the system or there exists on-site or off-diagonal disorder in the system which can produce an energy-dependent terms in the reduced Hamiltonian. In this paper we only focus on the present geometry and the multichannel and disordered cases will be considered in separated ones. The transmission coefficient in such a one-channel structure can be calculated by the standard transfer matrix technique. The transfer matrix can be defined as follows \begin{eqnarray} \left(\begin{array}{c}\psi_{i+1}\\ \psi_{i}\end{array}\right)= & \left(\begin{array}{cc}Em^{-\frac{|\alpha_N(i)-\alpha_N(i+1)|}{2}} & -m ^{\frac{|\alpha_N(i)-\alpha_N(i-1)|-|\alpha_N(i)-\alpha_N(i+1)|}{2}} \\ 1 & 0 \end{array}\right) \left(\begin{array}{c}\psi_{i}\\ \psi_{i-1}\end{array} \right) \nonumber \\ & \equiv B_N(i) \left(\begin{array}{c}\psi_{i}\\ \psi_{i-1}\end{array}\right). \end{eqnarray} So along the 1D chain with stage number $N$, the wave function of an electron in two leads are related to each other via a total transfer matrix \begin{equation} \left(\begin{array}{c}\psi_{n^N+1}\\ \psi_{n^N}\end{array}\right)= \prod_{i=0}^{n^N}B_N(n^N-i)\left(\begin{array}{c} \psi_0\\ \psi_{-1}\end{array}\right) \equiv {\bf T}(N) \left(\begin{array}{c}\psi_0\\ \psi_{-1} \end{array}\right). \end{equation} We demonstrate such an effective 1D chain structure in Fig. 2. This is a one-to-one mapping from the original network of Fig.1, and the resultant 1D chain shares the hierarchical structure at its longitudinal direction and the mirror reflecting symmetry as well. The self-similar features of the original structure are also kept in the mapping. The $\rho_N(i)$-structure in the longitudinal direction of the original system can be described by the following series: \begin{equation} 1,\underbrace{m^N,\cdots,m^N}_{n-1},\underbrace{m^{N-1}, \underbrace{m^N,\cdots,m^N}_{n-1},\cdots, m^N,\underbrace{m^{N-1}, \cdots,m^N}_{n-1}}_{n-1},m^{N-2},\cdots . \end{equation} One can see that this is a embedded multi-period structure. As a consequence the hoppings in the effective 1D chain are arranged as in the following series: \begin{equation} 1,m^{\frac{N}{2}}\underbrace{1,\cdots,1}_{n-2},m^{\frac{1}{2}}, \underbrace{m^{\frac{1}{2}}, \underbrace{1,\cdots,1}_{n-2},m^{\frac{1}{2}},\cdots, m^{\frac{1}{2}}, \underbrace{1,\cdots,1}_{n-2},m^{\frac{1}{2}}}_{n-2}, m^{\frac{1}{2}}, \underbrace{1,\cdots,1}_{n-2},m^{\frac{2}{2}},\cdots . \end{equation} Thus, one has the following recursion relation of the transfer matrix for $n\geq 2$: \begin{eqnarray} {\bf T}(N)= & B_N(n^N)B_N(n^N-1)B_{N-1}(n^{N-1}-1){\bf C}(N-1) [{\bf C}(N-1) \nonumber \\ & \times {\bf F}(N-1)]^{n-2}{\bf C}(N-1)B_{N-1}^{-1}(1)B_N(1)B_N(0), \end{eqnarray} where \[ {\bf C}(N-1)=B_{N-1}^{-1}(n^{N-1}){\bf T}(N-1)B_{N-1}^{-1}(0), \] and \[ {\bf F}(N-1)=\left( \begin{array}{cc} \frac{E}{\sqrt{m^{N-1}}} & -1 \\ 1 & 0 \end{array} \right) . \] One can find that the effective 1D chain contains $n^N$ bonds among which there are $(n-2)n^{N-1}$ bonds of hopping 1, 2 bonds of hopping $\sqrt{m^N}$, and $2n^{N-i-1}(n-1)$ bonds of hoppings $\sqrt{m^i}$ for $i=1,2,\cdots,N-1$. These bonds are arranged in a special aperiodic way so that one may expect a self-similar and Cantor-set-like transmission spectrum. From the transfer matrix {\bf T}, we can easily calculate the transmission coefficient for electrons from the left to the right lead: \begin{equation} \tau\left(E\right)=\frac{4\sin^2k}{[T_{21}-T_{12}+(T_{22}-T_{11})\cos k]^2+ (T_{11}+T_{22})^2 \sin^2k}, \end{equation} where $k=\arccos (E/2)$ is the wave vector of wave function in the leads. Thus, the conductance of the original network at the longitudinal direction is obtained according to Landauer formula \begin{equation} \label{lan} g(E)=\frac{e^2}{h}\int \tau(E) \frac {\partial f}{\partial E}dE \end{equation} where {\it f} is the Fermi distribution function \begin{equation} f(E,T)=\frac{1}{1+\exp{(E-E_f)/k_BT}}. \end{equation} \section{Transmission spectrum and temperature dependence of conductance} By using the above formulas we calculate the transmission coefficient as a function of energy for different structure parameters $(m,n)$ and different stage number. Typical results of transmission are illustrated in Fig. 3 and Fig. 4. We can see that when the stage number of the structure is low (say $N\leq 2$), $\tau (E)$ exhibits quite continuous variation with broad peaks and dips. However, when the stage number $N\geq 3$ the whole energy interval [-2,2] is divided into ``subbands'' which allow electron transmission and ``gaps'' of nearly zero transmission with sharp edges. By increasing $N$, the subbands are further divided into ``minibands'' and ``minigaps'', but the edges of the main subbands are almost independent of $N$. For large $N$, there are a large number of sparsely distributed peaks with $\tau(E)$ equal to or near unity which indicates almost complete transition corresponding to unscattered electronic states. This structure of the transmission spectrum also shows the self-similarity and Cantor-set-like behavior, like the situation in the Fibbonaci series and other quasiperiodic or aperiodic systems. By the comparison between Fig. 3 and Fig. 4, we find that although the number of the main subbands is the same but the width of the central main subband is slightly widened and the other two are shifted away from the band center by increasing $n$, which changes the number of the repeated bonds in the series. In the inset of Fig. 1(a) we also show the $N$ dependence of the transmission coefficient for energy $E=1.57$ which lies in the main band. It can be seen that for this energy by increasing $N$ the transmission first decreases but then increases. This behavior is sensitive to the position of $E$ as can be seen from the complicated structure in the transmission spectrum for large $N$. For the polymer networks, an interesting problem is to investigate the scaling behavior and the temperature dependence of the dc conductance. Since the transmission subbands exhibit ``dilute'' characteristics with a lot of transmission holes within them, the temperature dependence of the conductance is sensitive to the position of the Fermi level. From Eq. (\ref{lan}), for low temperature the conductance as a function of the Fermi level shows almost the same characteristics as those for the transmission coefficient as a function of energy. From Figs. 3 and 4 one can see that there exist energy points for which the transmission of large system is greater than that of small system. As a consequence for the system with Fermi level at these points the conductance increases with the system size provided that the temperature is low enough. This may be regarded as the ``metallic'' scaling behavior. There are still many other energy points for which the scaling behavior of the conductance is ``insulating'' as it decreases with the system size. However, the energy regimes for ``metallic'' and ``insulating'' states are strongly mixed with each other and form a complicated Cantor-set-like structure. If the temperature becomes higher, such a structure is smoothened by the thermal averaging. In Figs. 5 and 6 we plot the Fermi level dependence of the conductance at temperature $T=400$K for different structures and different system size. It can be seen that at this temperature the curves become smooth and in the whole energy range the conductance decreases with the increase of the system size, showing the insulating behavior. This is because more transmission holes appear in the transmission subbands when $N$ increases and this reduces the average of the transmission. In Fig. 7 we plot the temperature dependence of conductance $g(T)$ for one structure but with different Fermi levels. The curves manifest that the conductance can increase or decrease with the decrease of the temperature, strongly depending on the Fermi level $E_f$. The orders of magnitude of the variation of conductance at the range of [50K, 400K] could vary from 1 to 4. Only for carefully chosen values of $E_f$ (usually near edges of subbands for a given structure $(m,n))$, one could get a curve of $\log g(T)-T^{-1/2}$ whose slope can be fitted to the experimental data \cite{s1,s8}. In fact, the realistic stretched polymers should contain tremendous number of such networks in different stages, and doping might adjust Fermi level as well as the distribution of $(m,n)$. This may cause the deviation of the conductance calculated results in this simple model from the observed data. Nevertheless, as one can see from Fig. 7, our calculation displays a curve of good fitting to expression $\ln g_{dc} \propto T^{-1/2}$ within quite large range of temperature, while the solution of the model from the critical percolation approach in Ref. \cite{s8} demonstrates that $ g_{dc}\propto \frac{e^{-(T_0/T)^{1/2}}}{T^2}$ which slightly diverges from experimental results. \section{Conclusions and Discussion} In summary, we have shown that the hierarchical fractal network of Fig. 1 is equivalent to a 1D aperiodic series if only one-channel transport properties are under consideration. This equivalence is derived by summing over the transverse degree of freedom and, as a consequence, does not imply the one-dimensionality of this structure. The self-similarity and the aperiodicity of the original network are kept in this mapping. We calculate the transmission spectrum and temperature dependence of the conductance for systems with different structure parameters and various size. The transmission spectrum manifests Cantor-set-like characteristics, with complicated structures made of mixed transmission peaks and gaps, like most of the 1D quasiperiodic and aperiodic systems. Such a structure is also reflected in the mixed insulating and metallic scaling behavior of the conductance at low temperature. At higher temperature such a structure in the conductance is smoothened by the thermal averaging and the scaling behavior is insulator like in the whole energy range. This implies some type of metal-insulator transition at special values of the Fermi level, in agreement with the analysis in Ref. \cite{s9}. The magnitude and temperature dependence of the conductance strongly depend on the position of the Fermi level. By adjusting the value of $E_f$, one can reproduce the temperature dependence observed in the experiments. This simplified model can catch several features of polymer qualitatively but a complete description needs more sophisticated models. \section*{Acknowledgments} This work is supported by National Fund of Natural Science of China.